lectures on loop quantum gravity - arxiv · arxiv:gr-qc/0210094v1 28 oct 2002 lectures on loop...

92
arXiv:gr-qc/0210094v1 28 Oct 2002 Lectures on Loop Quantum Gravity T. Thiemann 1 MPI f. Gravitationsphysik, Albert-Einstein-Institut, Am M¨ uhlenberg 1, 14476 Golm near Potsdam, Germany Preprint AEI-2002-087 1 [email protected]

Upload: trinhdan

Post on 25-Jun-2018

228 views

Category:

Documents


0 download

TRANSCRIPT

arX

iv:g

r-qc

/021

0094

v1 2

8 O

ct 2

002

Lectures on Loop Quantum Gravity

T. Thiemann1

MPI f. Gravitationsphysik, Albert-Einstein-Institut,Am Muhlenberg 1, 14476 Golm near Potsdam, Germany

Preprint AEI-2002-087

[email protected]

Abstract

Quantum General Relativity (QGR), sometimes called Loop Quantum Gravity, has matured overthe past fifteen years to a mathematically rigorous candidate quantum field theory of the gravita-tional field. The features that distinguish it from other quantum gravity theories are 1) backgroundindependence and 2) minimality of structures.

Background independence means that this is a non-perturbative approach in which one does notperturb around a given, distinguished, classical background metric, rather arbitrary fluctuations areallowed, thus precisely encoding the quantum version of Einstein’s radical perception that gravity isgeometry.

Minimality here means that one explores the logical consequences of bringing together the twofundamental principles of modern physics, namely general covariance and quantum theory, withoutadding any experimentally unverified additional structures such as extra dimensions, extra symme-tries or extra particle content beyond the standard model. While this is a very conservative approachand thus maybe not very attractive to many researchers, it has the advantage that pushing the theoryto its logical frontiers will undoubtedly either result in a successful theory or derive exactly whichextra structures are required, if necessary. Or put even more radically, it may show which basicprinciples of physics have to be given up and must be replaced by more fundamental ones.

QGR therefore is, by definition, not a unified theory of all interactions in the standard sensesince such a theory would require a new symmetry principle. However, it unifies all presently knowninteractions in a new sense by quantum mechanically implementing their common symmetry group,the four-dimensional diffeomorphism group, which is almost completely broken in perturbative ap-proaches.

Contents

I Motivation and Introduction 5I.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

I.1.1 Why Quantum Gravity in the 21’stCentury ? . . . . . . . . . . . . . . . . . . . . . . . 6I.1.2 The Role of Background Independence . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

I.2 Introduction: Classical Canonical Formulation of General Relativity . . . . . . . . . . . . . . 15I.2.1 The ADM Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16I.2.2 Gauge Theory Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

I.3 Canonical Quantization Programme for Theories with Constraints . . . . . . . . . . . . . . . 22I.3.1 Refined Algebraic Quantization (RAQ) . . . . . . . . . . . . . . . . . . . . . . . . . . 22I.3.2 Selected Examples with First Class Constraints . . . . . . . . . . . . . . . . . . . . . . 25

II Mathematical and Physical Foundations of Quantum General Rel-ativity 28

II.1 Mathematical Foundations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29II.1.1 Polarization and Preferred Poisson Algebra B . . . . . . . . . . . . . . . . . . . . . . . 29II.1.2 Representation Theory of B and Suitable Kinematical Representations . . . . . . . . . 33

II.1.2.1 Curves, Paths, Graphs and Groupoids . . . . . . . . . . . . . . . . . . . . . . 33II.1.2.2 Topology on A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35II.1.2.3 Measures on A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36II.1.2.4 Representation Property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

II.2 Quantum Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40II.2.1 The Space of Solutions to the Gauss and Spatial Diffeomorphism Constraint . . . . . 41II.2.2 Kinematical Geometrical Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

III Selected Areas of Current Research 45III.1 Quantum Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

III.1.1 A Possible New Mechanism for Avoiding UV Singularities in Background IndependentQuantum Field Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

III.1.2 Sketch of a Possible Quantization of the Hamiltonian Constraint . . . . . . . . . . . . 50III.2 Loop Quantum Cosmology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

III.2.1 A New Approach To Quantum Cosmology . . . . . . . . . . . . . . . . . . . . . . . . . 56III.2.2 Spectacular Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

III.3 Path Integral Formulation: Spin Foam Models . . . . . . . . . . . . . . . . . . . . . . . . . . 58III.3.1 Spin Foams from the Canonical Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 58III.3.2 Spin Foams and BF – Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

III.4 Quantum Black Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62III.4.1 Isolated Horizons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62III.4.2 Entropy Counting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

1

III.5 Semiclassical Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67III.5.1 The Complexifier Machinery for Generating Coherent States . . . . . . . . . . . . . . 67III.5.2 Application to QGR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

III.6 Gravitons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75III.6.1 The Isomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75III.6.2 Induced Fock Representation With Polymer – Excitations . . . . . . . . . . . . . . . . 77

IV Selection of Open Research Problems 80

2

This is the expanded version of a talk given by the author at the 271st WE Heraeus Seminar “Aspectsof Quantum Gravity: From Theory to Experimental Search”, Bad Honnef, Germany, February 25th– March 1st 2002, http://www.uni-duesseldorf.de/QG-2002.

Basically, we summarize the present status of Canonical Quantum General Relativity (QGR),also known as “Loop Quantum Gravity”. Our presentation tries to be precise and at the same timetechnically not too complicated by skipping the proofs of all the statements made. These manymissing details, which are relevant to the serious reader, can be found in the notation used here inthis overview e.g. in the recent, close to exhaustive review [1] and references therein. Of course, inorder to be useful as a text for first reading we did not include all the relevant references here. Weapologize for that to the researchers in the field but we hope that a close to complete list of theirwork can be found in [1]. Nice reports, treating complementary subjects of the field and more generalaspects of quantum gravity can be found in [2].

The text is supplemented by numerous exercises of varying degree of difficulty whose purpose isto cut the length of the exposition and to arouse interest in further studies. Solving the problemsis not at all mandatory for an undertsanding of the material, however, the exercises contain furtherinformation and thus should be looked at even on a first reading.

On the other hand, if one solves the problems then one should get a fairly good insight into thetechniques that are important in QGR and in principle could serve as a preparation for a diplomathesis or a dissertation in this field. The problems sometimes involve mathematics that may beunfamiliar to students, however, this should not scare off but rather encourage the serious reader tolearn the necessary mathematical background material. Here is a small list of mathematical texts,from the author’s own favourites, geared at theoretical and mathematical physicists, that might behelpful:

• GeneralA fairly good encyclopedia isY. Choquet-Bruhat, C. DeWitt-Morette, “Analysis, Manifolds and Physics”, North Holland,Amsterdam, 1989

• General TopologyA nice text, adopting almost no prior knowledge isJ. R. Munkres, “Toplogy: A First Course”, Prentice Hall Inc., Englewood Cliffs (NJ), 1980

• Differential and Algebraic GeometryA modern exposition of this classical material can be found inM. Nakahara, “Geometry, Topology and Physics”, Institute of Physics Publishing, Bristol, 1998

• Functional AnalysisThe number one, unbeatable and close to complete exposition isM. Reed, B. Simon, “Methods of Modern Mathematical Physics”, vol. 1 – 4, Academic Press,New York, 1978especially volumes one and two.

• Measure TheoryAn elementary introduction to measure theory can be found in the beautiful bookW. Rudin, “Real and Complex Analysis”, McGraw-Hill, New York, 1987

3

• Operator AlgebrasAlthough we do not really make use of C∗−algebras in this review, we hint at the importanceof the subject, so let us includeO. Bratteli, D. W. Robinson, “Operator Algebras and Quantum Statistical Mechanics”, vol.1,2, Springer Verlag, Berlin, 1997

• Harmonic Analysis on GroupsAlthough a bit old, it still contains a nice collection of the material around the Peter & Weyltheorem:N. J. Vilenkin, “Special Functions and the Theory of Group Representations”, American Math-ematical Society, Providence, Rhode Island, 1968

• Mathematical General RelativityThe two leading texts on this subject areR. M. Wald, “General Relativity”, The University of Chicago Press, Chicago, 1989S. Hawking, Ellis, “The Large Scale Structure of Spacetime”, Cambridge University Press,Cambridge, 1989

• Mathematical and Physical Foundations of Ordinary QFTThe most popular books on axiomatic, algebraic and constructive quantum field theory areR. F. Streater, A. S. Wightman, “PCT, Spin and Statistics, and all that”, Benjamin, NewYork, 1964R. Haag, “Local Quantum Physics”, 2nd ed., Springer Verlag, Berlin, 1996J. Glimm, A. Jaffe, “Quantum Physics”, Springer-Verlag, New York, 1987

In the first part we motivate the particular approach to a quantum theory of gravity, called (Canon-ical) Quantum General Relativity, and develop the classical foundations of the theory as well as thegoals of the quantization programme.

In the second part we list the solid results that have been obtained so far within QGR. Thus, wewill apply step by step the quantization programme outlined at the end of section I.3 to the classicaltheory that we defined in section I.2. Up to now, these steps have been completed approximatelyuntil step vii) at least with respect to the Gauss – and the spatial diffeomorphism constraint. Theanalysis of the Hamiltonian constraint has also reached level vii) already, however, its classical limitis presently under little control which is why we discuss it in part three where current research topicsare listed.

In the third part we discuss a selected number of active research areas. The topics that we willdescribe already have produced a large number of promising results, however, the analysis is in mostcases not even close to being complete and therefore the results are less robust than those that wehave obtained in the previous part.

Finally, in the fourth part we summarize and list the most important open problems that wefaced during the discussion in this report.

4

Part I

Motivation and Introduction

5

I.1 Motivation

I.1.1 Why Quantum Gravity in the 21’stCentury ?

Students that plan to get involved in quantum gravity research should be aware of the fact that inour days, when financial resources for fundamental research are more and more cut and/or more andmore absorbed by research that leads to practical apllications on short time scales, one should have agood justification for why tax payers should support any quantum gravity research at all. This seemsto be difficult at first due to the fact that even at CERN’s LHC we will be able to reach energies ofat most 104 GeV which is fifteen orders of magnitude below the Planck scale which is theenergy scale at which quantum gravity is believed to become important. Therefore one could arguethat quantum gravity research in the 21’st century is of purely academic interest only.

To be sure, it is a shame that one has to justify fundamental research at all, a situation unheardof in the beginning of the 20’th century which probably was part of the reason for why so manybreakthroughs especially in fundamental physics have happened in that time. Fundamental researchcan work only in absence of any pressure to produce (mainstream) results, otherwise new, radical andindependent thoughts are no longer produced. To see the time scale on which fundamental researchleads to practical results, one has to be aware that General Relativity (GR) and Quantum Theory(QT) were discovered in the 20’s and 30’s already but it took some 70 years before quantum mechanicsthrough, e.g. computers, mobile phones, the internet, electronic devices or general relativity throughe.g. space travel or the global positioning system (GPS) became an integral part of life of a largefraction of the human population. Where would we be today if the independent thinkers of thosetimes were forced to do practical physics due to lack of funding for analyzing their fundamentalquestions ?

Of course, in the beginning of the 20’th century, one could say that physics had come to somesort of crisis, so that there was urgent need for some revision of the fundamental concepts: ClassicalNewtonian mechanics, classical electrodynamics and thermodynamics were so well understood thatMax Planck himself was advised not to study physics but engineering. However, although from apractical point of view all seemed well, there were subtle inconsistencies among these theories if onedrove them to their logical frontiers. We mention only three of them:1) Although the existence of atoms was by far not widely accepted at the end of the 19th century(even Max Planck denied them), if they existed then there was a serious flaw, namely, how shouldatoms be stable ? Acceralated charges radiate Bremsstrahlung according to Maxwell’s theory, thusan electron should fall into the nucleus after a finite amount of time.2) If Newton’s theory of absolute space and time was correct then the speed of light should depend onthe speed of the inertial observer. The fact that such velocity dependence was ruled out to quadraticorder in v/c in the famous Michelson-Morley experiment was explained by postulating an unknownmedium, called ether, with increasingly (as measurement precision was refined) bizarre properties inorder to conspire to a negative outcome of the interferometer experiment and to preserve Newton’snotion of space and time.3) The precession of mercury around the sun contradicted the ellipses that were predicted by Newton’stheory of gravitation.

Today we easily resolve these problems by 1) quantum mechanics, 2) special relativity and 3)general relativity. Quantum mechanics does not allow for continuous radiation but predicts a discreteenergy spectrum of the atom, special relativity removed the absolute notion of space and time andgeneral relativity generalizes the static Minkowski metric underlying special relativity to a dynamicaltheory of a metric field which revolutionizes our understanding of gravity not as a force but as

6

geometry. Geometry is curved at each point in a manifold proportional to the matter density atthat point and in turn curvature tells matter what are the straightest lines (geodesics) along whichto move. The ether became completely unnecessary by changing the foundation of physics andbeautifully demonstrates that driving a theory to its logical frontiers can make extra structuresredundant, what one had to change is the basic principles of physics.1

This historic digression brings us back to the motivation for studying quantum gravity in thebeginning of the 21st century. The question is whether fundamental physics also today is in a kindof crisis. We will argue that indeed we are in a situation not unsimilar to that of the beginning ofthe 20th century:Today we have very successful theories of all interactions. Gravitation is described by general relativ-ity, matter interactions by the standard model of elementary particle physics. As classical theories,their dynamics is summarized in the classical Einstein equations. However, there are several problemswith these theories, some of which we list below:

i) Classical – Quantum InconsistencyThe fundamental principles collide in the classical Einstein equations

Rµν − 1

2Rgµν

︸ ︷︷ ︸Geometry (GR, gen. covariance)

= κ Tµν(g)︸ ︷︷ ︸Matter (Stand.model, QT)

These equations relate matter density in form of the energy momentum tensor Tµν and geom-etry in form of the Ricci curvature tensor Rµν . Notice that the metric tensor gµν enters alsothe definition of the energy momentum tensor. However, while the left hand side is describeduntil today only by a classical theory, the right hand side is governed by a quantum fieldtheory (QFT). Since complex valued functions and operators on a Hilbert space are two com-pletely different mathematical objects, the only way to make sense out of the above equationswhile keeping the classical and quantum nature of geometry and matter respectively is to takeexpectation values of the right hand side, that is,

Rµν −1

2Rgµν = κ < Tµν(g0) >, κ = 8πGNewton/c

3

Here g0 is an arbitrary background metric, say the Minkowski metric η = diag(−1, 1, 1, 1).However, even if the state with respect to which the expectation value is taken is the vacuumstate ψg0 with respect to g0 (the notion of vacuum depends on the background metric itself,see below), the right hand side is generically non-vanishing due to the vacuum fluctuations,enforcing g = g1 6= g0. Hence, in order to make this system of equations consistent, onecould iterate the procedure by computing the vacuum state ψg1 and reinserting g1 into Tµν(.),resulting in g2 6= g1 etc. hoping that the procedure converges. However, this is generically notthe case and results in “run – away solutions” [3].

Hence, we are enforced to quantize the metric itself, that is, we need a quantum theory of

1 Notice, that the stability of atoms is still not satisfactorily understood even today because the full problem alsotreats the radiation field, the nucleus and the electron as quantum objects which ultimately results in a problem inQED, QFD and QCD for which we have no entirely satisfactory description today, see below.

7

gravity resulting in the

′′

Rµν −

1

2Rgµν = κ Tµν(g)

Quantum - Einstein - Equations

′′

The inverted commas in this equation are to indicate that this equation is to be made rigorousin a Hilbert space context. QGR is designed to exactly do that, see section III.1.

ii) General Relativity InconsistenciesIt is well-known that classical general relativity is an incomplete theory because it predictsthe existence of so-called spacetime singularities, regions in spacetime where the curvature orequivalently the matter density becomes infinite [4]. The most prominent singularities of thiskind are black hole and big bang singularities and such singularities are generic as shown inthe singularity theorems due to Hawking and Penrose. When a singularity appears it meansthat the theory has been pushed beyond its domain of validity, certainly when matter collapsesit reaches a state of extreme energy density at which quantum effects become important. Aquantum theory of gravity could be able to avoid these singularities in a similar way as quantummechanics explains the stability of atoms. We will see that QGR is able to achieve this, atleast in the simplified context of “Loop Quantum Cosmology”, see section III.2.

iii) Quantum Field Theory InconsistenciesAs is well-known, QFT is plagued by UV (or short distance) divergences. The fundamentaloperators of the theory are actually not operators but rather operator valued distributions andusually interesting objects of the theory are (integrals of) polynomials of those evaluated in thesame point. However, the product of distributions is, by definition, ill-defined. The appearanceof these divergences is therefore, on the one hand, not surprising, on the other hand it indicatesagain that the theory is incomplete: In a complete theory there is no room for infinities. Thus,either the appropriate mathematical framework has not been found yet, or they arise becauseone neglected the interaction with the gravitational field. In fact, in renormalizable theoriesone can deal with these infinities by renormalization, that is, one introduces a short distancecut-off (e.g. by point splitting the operator valued distributions) and then redefines massesand coupling constants of the theory in a cut-off dependent way such that they stay finite asthe cut-off is sent to zero. This redefinition is done in the framework of perturbation theory(Feynman diagrammes) by subtracting counter terms from the original Lagrangean which areformally infinite and a theory is said to be renormalizable if the number of algebraically differentsuch counter terms is finite.

The occurance of UV singularities is in deep conflict with general relativity due to the followingreason: In perturbation theory, the divergences have their origin in Feynman loop integralsin momentum space where the inner loop 4-momentum k = (E, P ) can become arbitrarilylarge, see figure 1 for an example from QED (mass renormalization). Now such virtual (off-shell) particles with energy E and momentum P have a spatial extension of the order of theCompton radius λ = ~/P and a mass of the order of E/c2. Classical general relativity predictsthat this lump of energy turns into a black hole once λ reaches the Schwarzschild radius of

8

p − k

p k p

Figure 1: One loop correction to the electron propagator in QED

the order of r = GE/c4. In a Lorentz frame where E ≈ Pc this occurs at the Planck energyE = EP =

√~/κc ≈ 1019GeV or at the Planck length Compton radius ℓP =

√~κ ≈ 10−33cm.

However, when a (virtual) particle turns into a black hole it completely changes its properties.For instance, if the virtual particle is an electron then it is able to interact only electroweaklyand thus can radiate only particles of the electrowak theory. However, once a black hole hasformed, also Hawking processes are possible and now any kind of particles can be emitted, but ata different production rate. Of course, this is again an energy regime at which quantum gravitymust be important and these qualitative pictures must be fundamnetally wrong, however, theyshow that there is a problem with integrating virtual loops into the UV regime. In fact, thesequalitative thoughts suggest that gravity could serve as a natural cut-off because a black holeof Planck mass size ℓP should decay within a Planck time unit tP = ℓP/c ≈ 10−43s so thatone has to integrate P only until EP/c. Moreover, it indicates that spacetime geometry itselfacquires possibly a discrete structure since arguments of this kind make it plausible that itis impossible to resolve spacetime distances smaller than ℓP , basically because the spacetimebehind an event horizon is in some sense “invisible”. These are, of course, only hopes and mustbe demonstrated within a concrete theory. We will see that QGR is able to precisely do thatand its fundamental discreteness is in particular responsible for why the Bekenstein Hawkingentropy of black holes is finite, see sections II.2.2, III.1.1 and III.4.

So we see that there is indeed a fundamental inconsisteny within the current description of fun-damental physics comparable to the time before the discovery of GR and QT and its resolution,Quantum Gravity, will revolutionize not only our understanding of nature but will also drive newkinds of technology that we do not even dare to dream of today.

I.1.2 The Role of Background Independence

Given the fact that both QT and GR were discovered already more than 70 years ago and that peoplehave certainly thought about quantizing GR since then and that matter interactions are more or lesssuccesfully described by ordinary quantum field theories (QFT), it is somewhat surprising that wedo not yet have a quantum gravity theory. Why is it so much harder to combine gravity with theprinciples of quantum mechanics than for the other interactions ? The short answer is that

9

supp(f’)

supp(f)

Figure 2: Spacelike separated regions in Minkowski space

Ordinary QFT only incorporates Special Relativity.

To see why, we just have to remember that ordinary QFT has an axiomatic definition, here fora scalar field for simplicity:

WIGHTMAN AXIOMS (Scalar Fields on Minkowsk Space)

W1 Poincare Group P:∃ continuous, unitary representation U of P on a Hilbert space H.

W2 Forward Lightcone Spektral Condition:For the generators P µ of the translation subgroup of P holds ηµνP

µP ν ≤ 0, P 0 ≥ 0.

W3 Existence and Uniqueness of a P−inariant Vacuum Ω:U(p)Ω = Ω ∀p ∈ P.

W4 P−Covariance:

φ(f) :=

∫dD+1xf(x)φ(x), f ∈ S(RD+1)

φ(f1)..φ(fn)Ω dense in H and U(p)φ(f)U(p)−1 = φ(f p)

W5 Locality (Causality):If supp(f), supp(f ′) spacelike separated (see figure 2), then [φ(f), φ(f ′)] = 0.

It is obvious that due to the presence of the Minkowski backgrounde metric η we have available alarge amount of structure which forms the fundament on which ordinary QFT is built. Roughly, we

10

have the following scheme:

ηµν ⇒ P ⇒ H = P 0

b.-metric symm.-group Ham.-operator

⇓ ⇓

(x− y)2 = 0 ⇒ commutation Ωlightcone relations ground state

Notice that a generic background metric has no symmetry group at all so that it is not straightfor-ward to generalize these axioms to QFT on general curved backgrounds, however, since any metricis locally diffeomorphic to the Minkowski metric, a local generalization is possible and results in theso-called microlocal analysis in which the role of vacuum states is played by Hadamard states, seee.g. [5].

The fundamental, radically new feature of Einstein’s theory is that thereis no background metric at all: The theory is background independent.The lightcones themselves are fluctuating, causality and locality becomeempty notions. The dome of ordinary QFT completely collapses.

Of course, there must be a regime in any quantum gravity theory where the quantum fluctuations ofthe metric operator are so tiny that we recover the well established theory of free ordinary quantumfields on a given background metric, however, the large fluctuations of the metric operator can nolonger be ignored in extreme astrophysical or cosmological situations, such as near a black hole orbig bang singularity.

People have tried to rescue the framework of ordinary QFT by splitting the metric into a back-ground piece and a fluctuation piece

gµν = ηµν + hµν

↑ ↑ ↑full metric background (Minkowski) perturbation (graviton)

(I.1.2.1)

which results in a Lagrangean for the graviton field hµν and could in principle be the definition of agraviton QFT on Minkowski space. However, there are serious drawbacks:

i) Non-RenormalizabilityThe resulting theory is perturbatively non-renormalizable [6] as could have been expected fromthe fact that the coupling constant of the theory, the Planck area ℓ2P , has negative mass di-mension (in Planck units). Even the supersymmetric extension of the theory, in any possibledimension has this bad feature [7]. It could be that the theory is non-perturbatively renormal-izable, meaning that it has a non-Gaussian fix point in the language of Wilson, a possibilitythat has recently regained interest [8].

ii) Violation of Background IndependenceThe split of the metric performed above again distinguishes the Minkowski metric among allothers and reintroduces therefore a background dependence. This violates the key featureof Einstein’s theory and thus somehow does not sound correct, we better keep background

11

independence if we want to understand how quantum mechanics can possibly work togetherwith general covariance.

iii) Violation of Diffeomorphism CovarianceThe split of the metric performed above is certainly not diffeomorphism covariant, it breaksthe diffeomorphism group down to Poincare group. Violation of fundamental, local gaugesymmetries is usually considered as a bad feature in Yang-Mills theories on which all the otherinteractions are based, thus already from this point of view perturbation theory looks dangerous.As a side remark we see that background dependence and violation of general covariance aresynonymous.

iv) Gravitons and GeometrySomehow the whole idea of the gravitational interaction as a result of graviton exchange on abackground metric contradicts Einstein’s original and fundamental idea that gravity is geometryand not a force in the usual sense. Therefore such a perturbative description of the theory isvery unnatural from the outset and can have at most a semi-classical meaning when the metricfluctuations are very tiny.

v) Gravitons and DynamicsAll that classical general relativity is about is how a metric evolves in time in an interplaywith the matter present. It is clear that an initially (almost) Minkowskian metric can evolve tosomething that is far from Minkowskian at other times, an example being cosmological big bangsituations or the collapse of initially dilluted matter (evolved backwards). In such situationsthe assumption being made in (I.1.2.1), namely that h is “small” as compared to η is just notdynamically stable. In some sense it is like trying to use Cartesian coordinates for a spherewhich can work at most locally.

All these points just naturally ask for a non-perturbative approach to quantum gravity. This, inturn, could also cure another unpleasant feature about ordinary QFT: Today we do not have a singleexample of a rigorously defined interacting ordinary QFT in four dimensions, in other words, therenormalizable theories that we have are only defined order by order in perturbation theory but theperturbation series diverges. A non-perturbative definition, to which we seem to be forced whencoupling gravity anyway, might change this unsatisfactory situation.

It should be noted here that there is in fact a consistent perturbative description of a candidatequantum gravity theory, called string theory (or M – Theory nowadays) [9].2 However, in orderto achieve this celebrated rather non-trivial result, expectedly one must introduce extra structure:The theory lives in 10 (or 11) rather than 4 dimensions, it is necessarily supersymmetric and it hasan infinite number of extra particles besides those that are needed to make the theory compatiblewith the standard model. Moreover, at least as presently understood, again the fundamental newingredient of Einstein’s theory, background independence, is violated in string theory. This currentbackground dependence of string theory is supposed to be overcome once M – Theory has beenrigorously defined.

At present only string theory has a chance to explain the matter content of our universe. Theunification of symmetries is a strong guiding principle in physics as well and has been pushed also

2String theory is an ordinary QFT but not in the usual sense: It is an ordinary scalar QFT on a 2d Minkowskispace, however, the scalar fields themselves are coordinates of the ambient target Minkowski space which in this caseis 10 dimensional. Thus, it is similar to a first quantized theory of point particles. The theory is renormalizable andpresumably even finite order by order in perturbation theory but the perturbation series does not converge.

12

Figure 3: QFT on Background Spacetime (M, g0): Actor: Matter, Stage: Geometry + ManifoldM .

by Einstein in his programme of geometrization of physics attemting to unify electromagnetism andgravity in a five-dimensional Kaluza – Klein theory. The unification of the electromagnetic and theweak force in the electroweak theory is a prime example for the success of such ideas. However,unification of forces is an additional principle completely independent of background independenceand is not necessarily what a quantum theory of gravity must achieve: Unification of forces can beanalyzed at the purely classical level3. Thus, the only question is whether the theory can be quan-tized before unification or not (should unification of geometry and matter be realized in nature at all).

We are therefore again in a situation, similar to that before the discovery of special relativity, wherewe have the choice between a) preserving an old principle, here renormalizability of perturbativeQFT on background spacetimes (M, η), at the price of introducing extra structure (extra unifica-tion symmetry), or b) replacing the old principle by a new principle, here non-perturbative QFTon a differentiable manifold M , without new hypothetical structure. At this point it unclear whichmethodology has more chances for success, historically there is evidence for either of them (e.g. theunification of electromagnetism and the massive Fermi model is evidence for the former, the replace-ment of Newton’s notion of spactime by special relativity is evidence for the latter) and it is quitepossible that we actually need both ideas. In QGR we take the latter point of view to begin withsince there maybe zillions of ways to unify forces and it is hard to judge whether there is a “naturalone”, therefore the approach is purposely conservative because we actually may be able to derive anatural way of unification, if necessary, if we drive the theory to its logical frontiers. Among thevarious non-perturbative approaches available we will choose the canonical one.

Pictorially, one could illustrate the deep difference between a background dependent QFT andbackground independent QFT as follows: In figure 3 we see matter in the form of QCD (notice thequark (Q) propagators, the quark-gluon vertices and the three – and four point gluon (G) vertices)displayed as an actor in green. Matter propagates on a fixed background spacetime g0 accordingto well-defined rules, particles know exactly what timelike geodesics are etc. This fixed background

3In fact, e.g. the unified electroweak SU(2)L×U(1) theory with its massless gauge bosons can be perfectly describedby a classical Lagrangean. The symmetry broken, massive U(1) theory can be derived from it, also classically, byintroducing a constant background Higgs field (Higgs mechanism) and expanding the symmetric Lagrangean aroundit. It is true that the search for a massless, symmetric theory was inspired by the fact that a theory with massive gaugebosons is not renormalizable (so the motivation comes from quantum theory) and, given the non-renormalizability ofgeneral relativity, many take this as an indication that one must unify gravity with matter, one incarnation of whichis string theory. However, the argument obviously fails should it be possible to quantize gravity non-perturbatively.

13

Figure 4: QGR on Differential Manifold M : Actor: Matter + Geometry, Stage: Manifold M

spacetime g0 is displayed as a firm stage in blue. This is the situation of a QFT on a BackgroundSpacetime.

In contrast, in figure 4 the stage has evaporated, it has become itself an actor (notice the arbitrarilyhigh valent graviton (g) vertices) displayed in blue as well. Both matter and geometry are nowdynamical entities and interact as displayed by the red vertex. There are no light cones any longer,rather the causal structure is a semiclassical concept only. This is the situation of a QFT on aDifferential Manifold and this is precisely what QGR aims to rigorously define.

It is clear from these figures that the passage from a QFT on a background spacetime to a QFTon a differential manifold is a very radical one: It is like removing the chair on which you sit andtrying to find a new, yet unknown, mechanism that keeps you from falling down. We should mentionhere that for many researchers in quantum gravity even that picture is not yet radical enough, someproposals require not only to get rid of the background metric g0 but also of the differential manifold,allowing for topology change. This is also very desired in QGR but considered as a second step. In3d QGR also this step could be completed and the final picture is completely combinatorial.

Let us finish this section by stating once more what we mean by Quantum General Relativity (QGR).

14

DEFINITION

(Canonical) Quantum General Relativity (QGR) is an attempt toconstruct a mathematically rigorous, non – perturbative, back-ground independent Quantum Field Theory of four – dimensional,Lorentzian General Relativity and all known matter in the contin-uum.

No additional, experimentally unverified structures are intro-duced. The fundamental principles of General Covariance andQuantum Theory are brought together and driven to their logicalfrontiers guided by mathematical consistency.

QGR is not a unified theory of all interactions in the stan-dard sense since unification of gauge symmetry groups is notnecessarily required in a non-perturbative approach. However,Geometry and Matter are unified in a non – standard sense bymaking them both transform covariantly under the DiffeomorphismGroup at the quantum level.

I.2 Introduction: Classical Canonical Formulation of Gen-

eral Relativity

In this section we sketch the classical Hamiltonian formulation of general relativity in terms ofAshtekar’s new variables. There are many ways to arrive at this new formulation and we will choosethe one that is the most convenient one for our purposes.

The Hamiltonian formulation by definition requires some kind of split ot the spacetime variablesinto time and spatial variables. This seems to contradict the whole idea of general covariance,however, quantum meachanics as presently formulated requires a notion of time because we interpretexpection values of operators as instantaneous measurement values averaged over a large numberof measurements. In order to avoid this one has to “covariantize” the interpretation of quantummechanics, in particular the measurement process, see e.g. [10] for a discussion. There are a numberof proposals to make the canonical formulation more covariant, e.g.4: Multisymplectic Ansatze [13] inwhich there are multimomenta, one for each spacetime dimension, rather than just one for the timecoordinate; Covariant phase space formulations [14] where one works on the space of solutions to thefield equations rather than on the intial value instantaneous phase space; Peierl’s bracket formulations[15] which covariantize the notion of the usual Poisson bracket; history bracket formulations [16],which grew out of the consistent history formulation of quantum mechanics [17], and which extends

4Path integrals [11] use the Lagrangean rather than the Hamiltonian and therefore seem to be better suitedto a covariant formulation than the canonical one, however, usually the path integral is interpreted as some sort ofpropagator which makes use of instantaneous time Hilbert spaces again which therefore cannot be completely discardedwith. At present, this connection with the canonical formulation is not very transparent, part of the reason being thatthe path integral is usually only defined in its Euclidean formulation, however the very notion of analytic continuationin time is not very meaningful in a theory where there is no distinguished choice of time, see however [12] for recentprogress in this direction.

15

the usual spatial Poisson bracket to spacetime.At the classical level all these formulations are equivalent. However, at the quantum level, one

presently gets farthest within the the standard canonical formulation: The quantization of the mul-tisymplectic approach is still in its beginning, see [18] for the most advanced results in this respect;The covariant phase space formulation is not only very implicit because one usually does not knowthe space of solutions to the classical field equations, but even if one manages to base a quantumtheory on it, it will be too close to the classical theory since certainly the singularities of the classicaltheory are also built into the quantum theory; The Peierl’s bracket also needs the explicit space ofsolutions to the classical field equations; Also the quantization of the history bracket formulationjust has started, see [19] for first steps in that direction.

Given this present status of affairs, we will therefore proceed with the standard canonical quanti-zation and see how far we get. Notice that there is no obvious problem with general covariance: Forinstance, standard Maxwell theory can be quantized canonically without any problem and one canshow that the theory is Lorentz covariant although the spacetime split into space and time seemsto break the Lorentz group down to the rotation group. This is not at all the case ! It is just thatLorentz covariance is not manifest, one has to do some work in order to establish Lorentz covariance.Indeed, as we will see, at least at the classical level we will explicitly recover the four-dimensionaldiffeomorphism group in the formalism, although it is admittedly deeply hiddeen in the canonicalformalism.

With these cautionary remarks out of the way, we will thus assume that the four dimensionalspacetime manifold has the topology R × σ, where σ is a three dimensional manifold of arbitrarytopology, in order to perform the 3 + 1 split. This assumption about the topology of M mayseem rather restrictive, however, it is not due to the following reasons: (1) According to a theoremdue to Geroch any globally hyperbolic manifold (roughly those that admit a smooth metric witheverywhere Lorentzian signature) are necessarily of that topology. Since Lorentzian metrics are whatwe are interested in, at least classically the assumption about the topology of M is forced on us.(2) Any four manifold M has the topology of a countable disjoint union ∪αIα × σα where either Iαare open intervals and σα is a three manifold or Iα is a one point set and σα is a two manifold (thelatter are the intersections of the closures of the former). In this most generic situation we thus allowtopology change between different three manifolds and it is even classically an open question how tomake this compatible with the action principle. We take here a practical point of view and try tounderstand the quantum theory first for a single copy of the form R × σ and later on worry how weglue the theories for different σ′s together.

I.2.1 The ADM Formulation

In this nice situation the 3+1 split is well known as the Arnowitt – Deser – Misner (ADM) formulationof general relativity, see e.g. [4] and we briefly sketch how this works. Since M is diffeomorphicto R × σ we know that M foliates into hypersurfaces Σt, t ∈ R as in figure 5, where t labelsthe hypersurface and will play the role of our time coordinate. If we denote the four dimensionalcoordinates by Xµ, µ = 0, 1, 2, 3 and the three dimensional coordinates by xa, a = 1, 2, 3 then weknow that there is a diffeomorphism ϕ : R × σ → M ; (t, x) 7→ X = ϕ(t, x) where Σt = ϕ(t, σ). Westress that the four diffeomorphism ϕ is completely arbitrary until this point and thus the foliationof M is not at all fixed, in other words, the set of foliations is in one to one correspondence withDiff(M), the four dimensional diffeomorphism group. Consider the tangential vector fields to Σt

given bySa(X) := (∂a)ϕ(t,x)=X = (ϕµ,a(t, x))ϕ(t,x)=X ∂µ (I.2.1.1)

16

Figure 5: Foliation of M

Denoting the four metric by gµν we define a normal vector field nµ(X) by gµνnµSνa = 0, gµνn

µnν = −1.Thus, while the tangential vector fields depend only on the foliation, the normal vector field dependsalso on the metric. Let us introduce the foliation vector field

T (X) := (∂t)ϕ(t,x)=X = (ϕµ,t(t, x))ϕ(t,x)=X ∂µ (I.2.1.2)

and let us decompose it into the basis n, Sa. This results in

T = Nn + UaSa (I.2.1.3)

where N is called the lapse function while UaSa is called the shift vector field. The arbitrariness of thefoliation is expressed in the arbitrariness of the fields N,Ua. We can now introduce two symmetricspacetime tensor fields (∇ is the unique, torsion free covariant differential compatible with gµν)

qµν = gµν + nµnν , Kµν = qµρqνσ∇ρnν (I.2.1.4)

called the intrinsic metric and the extrinsic curvature respectively which are spatial, that is, theircontraction with n vanishes. Thus, their full information is contained in their components withrespect to the spatial fields Sa, e.g. qab(t, x) = [qµνS

µaS

νb ](X(t, x)). In particular,

Kab(t, x) =1

2N[qab − LUqab] (I.2.1.5)

contains information about the velocity of qab. Here L the Lie derivative. The metric gµν is completelyspecified in terms of qab, N, U

a as one easily sees by expressing the line element ds2 = gµνdXµ dXν

in terms of dt, dxa.

Exercise I.2.1.Recall the definition of the Lie derivative and verify that Kµν is indeed symmetric and that formula(I.2.1.5) holds.Hint:A hypersurface Σt can be defined by the solution of an equation of the form τ(X) = t. Conclude thatnµ ∝ ∇µτ and use torsion – freness of ∇.

The Legendre transformation of the Einstein-Hilbert action

S =1

κ

M

d4X√

| det(g)|R(4) (I.2.1.6)

17

with qab, N, Ua considered as configuration coordinates in an infinite dimensional phase space is

standard and we will not repeat the analysis here, which uses the so - called Gauss - Codacciequations.

Here we are considering for simplicity only the case that σ is compact, otherwise (I.2.1.6) wouldcontain boundary terms. The end result is

S =1

κ

R

dt

σ

d3xqabP ab + NP + NaPa − [λP + λaPa + UaVa +NC] (I.2.1.7)

where

P ab =κδS

δqab=

√det(q)[qacqbd − qabqcd]Kcd (I.2.1.8)

and P, Pa are the momentua conjugate to qab, N, Ua respectively. Thus, we have for instance the

equal time Poisson brackets

P ab(t, x), P cd(t, y) = qab(t, x), qcd(t, y) = 0, P ab(t, x), qcd(t, y) = κδa(cδbd)δ(x, y) (I.2.1.9)

where (.)(ab) := [(.)ab + (.)ba]/2 denotes symmetrization. The functions C, Va which depend only onqab, P

ab are called the Hamiltonian and Spatial Diffeomorphism constraint respectively for reasonsthat will become obvious in a moment. Their explicit form is given by

Va = −2qacDbPbc

C =1√

det(q)[qacqbd −

1

2qabqcd]P

abP cd −√

det(q)R (I.2.1.10)

where D is the unique, torsion – free covariant differential compatible with qab and R is the curvaturescalar associated with qab.

The reason for the occurence of the Lagrange multipliers λ, λa is that the Lagrangean (I.2.1.6)is singular, that is, one cannot solve all the velocities in terms of momenta and therefore one mustuse Dirac’s procedure [20] for the Legendre transform of singular Lagrangeans. In this case thesingularity structure is such that the momenta conjugate to N,Ua vanish identically, whence theLagrange multipliers which when varied give the equations of motion P = Pa = 0. The equationsof motion with respect to the Hamiltonian (i.e. F := H,F for any functional F of the canonicalcoordinates)

H =

∫d3x[λP + λaPa + UaVa +NC] (I.2.1.11)

for N,Ua reveal that N,Ua are themselves Lagrange multipliers, i.e. completely unspecified functions(proportional to λ, λa) while the equations of motion for P, Pa give P = −C, Pa = −Va. Since P, Paare supposed to vanish, this requires C = Va = 0 as well. Thus we see that the Hamiltonian isconstrained to vanish in GR ! We will see that this is a direct consequence of the four dimensionaldiffeomorphism invariance of the theory.

Now the equations of motion for qab, Pab imply the so-called Dirac (or hypersurface deformation)

algebra

V (U), V (U ′) = κV (LUU ′)

V (U), C(N) = κC(LUN)

C(N), C(N ′) = κV (q−1(NdN ′ −N ′dN)) (I.2.1.12)

18

where e.g. C(N) =∫d3xNC. These equations tell us that the condition H = Va = 0 is preserved

under evolution, in other words, the evolution is consistent ! This is a non-trivial result. One says,the Hamiltonian and vector constraint form a first class constraint algebra. This algebra is much morecomplicated than the more familiar Kac-Moody algebras due to the fact that it is not an (infinite)dimensional Lie algebra in the true sense of the word because the “structure constants” on the righthand side of the last line in (I.2.1.12) are not really constants, they depend on the phase space. Suchalgebras are open in the the terminology of BRST [21] and about their representation theory onlyvery little is known.

Exercise I.2.2.Derive (I.2.1.12) from (I.2.1.9).Hint:Show first that the Poisson bracket between local functions which contain spatial derivatives is simplythe spatial derivatives applied to the Poisson bracket. Since the Poisson bracket of local functions isdistributional recall that derivatives of distributions are defined through an integration by parts.

Since the variables P, Pa drop out completely from the analysis and N,Ua are Lagrange multipli-ers, we may replace (I.2.1.7) by

S =1

κ

R

dt

σ

d3xqabP ab − [UaVa +NH ] (I.2.1.13)

with the understanding that N,Ua are now completely arbitrary functions which parameterize thefreedom in choosing the foliation. Since the Hamiltonian of GR depends on the completely unspeci-fied functions N,Ua, the motions that it generates in the phase space M coordinatized by (P ab, qab)subject to the Poisson brackets (I.2.1.9) are to be considered as pure gauge transformations. The in-finitesimal flow (or motion) of the canonical coordinates generated by the corresponding Hamiltonianvector fields on M has the following form for an arbitrary tensor tab built from qab, P

ab

V (U), tabEOM = κ(LU tab)C(N), tabEOM = κ(LNntab) (I.2.1.14)

where the subscript EOM means that these relations hold for generic functions on M only whenthe vacuum equations of motion (EOM) R

(4)µν − R(4)gµν/2 = 0 hold. Equation (I.2.1.14) reveals that

Diff(M) is implemented also in the canonical formalism, however, in a rather non-trivial way: Thegauge motions generated by the constraints can be interpreted as four-dimensional diffeomorphismsonly when the EOM hold. This was to be expected because a diffemorphism orthogonal to thehypersurface means evolution in the time parameter, what is surprising though is that this evolutionis considered as a gauge transformation in GR. Off the solutions, the constraints generate differentmotions, in other words, the set of gauge symmetries is not Diff(M) everywhere in the phase space.This is not unexpected: The action (I.2.1.6) is obviously Diff(M) invariant, but so would be any actionthat is an integral over a four-dimensional scalar density of weight one formed from polynomials in thecurvature tensor and its covariant derivatives. This symmetry is completely insensitive to the specificLagrangean in question, it is kinematical. The dynamics generated by a specific Lagrangean mustdepend on that Lagrangean, otherwise all Lagrangeans underlying four dimensionally diffeomorphisminvariant actions would equal each other up to a diffeomorphism which is certainly not the case(consider for instance higher derivative theories). In particular, that dynamics is, a priori, completelyindependent of Diff(M). As a consequence, Dirac observables, that is, functions on M which aregauge invariant (have vanishing Poisson brackets with the constraints), are not simply functionals of

19

Figure 6: Constraint hypersurface M and gauge orbit [m] of m ∈ M in M.

the four metric invariant under four diffeomorphisms because they must depend on the Lagrangean.The set of these dynamics dependent gauge transformations does not obviously form a group ashas been investigated by Bergmann and Komar [22]. The geometrical origin of the hypersurfacedeformation algebra has been investigated in [23]. Torre and Anderson have shown that for compactσ there are no Dirac observables which depend on only a finite number of spatial derivatives of thecanonical coordinates [24] which means that Dirac observables will be highly non-trivial to construct.

Let us summarize the gauge theory of GR in figure 6: The constraints C = Va = 0 define aconstraint hypersurface M within the full phase space M. The gauge motions are defined on allof M but they have the feature that they leave the constraint hypersurface invariant, and thusthe orbit of a point m in the hypersurface under gauge transformations will be a curve or gaugeorbit [m] entirely within it. The set of these curves defines the so-called reduced phase space andDirac observables restricted to M depend only on these orbits. Notice that as far as the counting isconcerned we have twelve phase space coordinates qab, P

ab to begin with. The four constraints C, Vacan be solved to eliminate four of those and there are still identifications under four independent setsof motions among the remaining eight variables leaving us with only four Dirac observables. Thecorresponding so-called reduced phase space has therefore precisely the two configuration degrees offreedom of general relativity.

I.2.2 Gauge Theory Formulation

We can now easily introduce the shift from the ADM variables qab, Pab to the connection variables

introduced first by Ashtekar [25] and later somewhat generalized by Immirzi [26] and Barbero [27].We introduce su(2) indices i, j, k, .. = 1, 2, 3 and co-triad variables eja with inverse eaj whose relationwith qab is given by

qab := δjkejaekb (I.2.2.1)

Defining the spin connection Γja through the equation

∂aejb − Γcabe

jc + ǫjklΓ

kaelb = 0 (I.2.2.2)

where Γcab are the Christoffel symbols associated with qab we now define

Aja = Γja + βKabebj , Ea

j =√

det(q)eaj/β (I.2.2.3)

20

where β ∈ C − 0 is called the Immirzi parameter. In this article we only consider real valued andpositive β. Finally we introduce the SU(2) Gauss constraint

Gj := ∂aEaj + ǫjklA

kaE

bl (I.2.2.4)

with ǫjkl the structure constants of su(2) which we would encounter in the canonical formulation ofany SU(2) gauge theory. As one can check, modulo Gj = 0 one can then write C, Va in terms of A,Eas follows

Va = F jabE

bj ,

C =F jabǫjklE

ajE

bl√

| det(E)|+ More (I.2.2.5)

where F = 2(dA + [A,A]) is the curvature of A and “More” is an additional term which is morecomplicated but can be treated by similar methods as the one displayed.

We then have the following theorem [25].

Theorem I.2.1.Consider the phase space M coordinatized by (Aja, E

bj ) with Poisson brackets

Eaj (x), E

bk(y) = Aja(x), Akb (y) = 0, Ea

j (x), Akb (y) = κδab δ

kj δ(x, y) (I.2.2.6)

and constraints Gj , C, Va. Then, solving only the constraint Gj = 0 and determining the Diracobservables with respect to it leads us back to the ADM phase space with constraints C, Va.

The proof of the theorem is non-trivial and tedious and can be found in the notation used herein [1]. Alternatively one can find directions for a proof in the subsequent exercise. In particular,this works only because the Gauss constraint is in evolution with itself and the other constraints,specifically G,G ∝ G, G, V = G,H = 0.

Exercise I.2.3.i)Prove theorem I.2.1.Hint:Express qab, P

ab in terms of Aja, Eaj by using (I.2.2.1), (I.2.2.2), (I.2.2.3), and (I.2.2.4) and check that

the Poisson brackets, with respect to (I.2.2.6), among the solutions qab = sab[A,E], P ab = Sab[A,E]equal precisely (I.2.1.9) modulo terms proportional to Gj.ii)Define G(Λ) :=

∫σd3xΛjGj and D(U) :=

∫σUa[Va − AjaGj ] and [Λ,Λ′]j = ǫjklΛ

k(Λ′)l. Verify thefollowing Poisson brackets

G(Λ), G(Λ′) = κG([Λ,Λ′])

G(Λ), V (U) = 0

D(U), D(U ′) = κD([U,U ′]) (I.2.2.7)

and conclude that the Hamiltonian vector fields of G(Λ) and D(U) respectively generate SU(2) gaugetransformations and spatial diffeomorphisms of σ respectively.

21

Hint:Show first that

G(Λ/κ), Aja(x) = −Λj,a + ǫjklΛ

kAla

D(U/κ), Aja(x) = U bAja,b + U b,aA

jb (I.2.2.8)

to conclude that A transforms as a connection under infinitesimal gauge transformations and as aone-form under infinitesimal diffeomorphisms. Consider then gt(x) := exp(tΛjτj/(2κ)) and ϕt(x) :=cU,x(t) where t 7→ cU,x(t) is the unique integral curve of U through x, that is, cU,x(t) = U(cU,x(t)), cU,x(0) =x. Recall that the usual transformation behaviour of connections and one-forms under finite gaugetransformations and diffeomorphisms respectively is given by (e.g. [28])

Ag = −dgg−1 + Adg(A)

Aϕ = ϕ∗A (I.2.2.9)

where A = Ajadxaτj/2, Adg(.) = g(.)g−1 denotes the adjoint representation of SU(2) on su(2) and ϕ∗

denotes the pull-back map of p−forms and iτj are the Pauli matrices so that τjτk = −δjk12 + ǫjklτl.Verify then that (I.2.2.8) is the derivative at t = 0 of (I.2.2.9) with g := gt, ϕ := ϕt. Similarly,derive that E transforms as an su(2)−valued vector field of density weight one. (Recall that a tensor

field t of some type is said to be of density weight r ∈ R if t√

| det(s)|−r is an ordinary tensor fieldof the same type where sab is any non-degenerate symmetric tensor field).

From the point of view of the classical theory we have made things more complicated: Insteadof twelve variables q, P we now have eighteen A,E. However, the additional six degrees of freedomare removed by the first class Gauss constraint which shows that working on our gauge theory phasespace is equivalent to working on the ADM phase space. The virtue of this extended phase spaceis that canonical GR can be formulated in the language of a canonical gauge theory where A playsthe role of an SU(2) connection with canonically conjugate electric field E. Besides the remark thatthis fact could be the starting point for a possible gauge group unification of all four forces we nowhave access to a huge arsenal of techniques that have been developed for the canonical quantizationof gauge theories. It is precisely this fact that has enabled steady progress in this field in the lastfifteen years while one was stuck with the ADM formulation for almost thirty years.

I.3 Canonical Quantization Programme for Theories with

Constraints

I.3.1 Refined Algebraic Quantization (RAQ)

As we have seen, GR can be formulated as a constrained Hamiltonian system with first class con-straints. The quantization of such systems has been considered first by Dirac [20] and was laterrefined by a number of authors. It is now known under the name refined algebraic quantization(RAQ). We will briefly sketch the main ideas following [29].

i) Phase Space and ConstraintsThe starting point is a phase space (M, ., .) together with a set of first class constraints CIand possibly a Hamiltonian H .

22

ii) Choice of PolarizationIn order to quantize the phase space we must choose a polarization, that is, a Lagrangeansubmanifold C of M which is called configuration space. The coordinates of C have vanishingPoisson brackets among themselves. If M is a cotangent bundle, that is, M = T ∗Q then itis natural to choose Q = C and we will assume this to be the case in what follows. For moregeneral cases, e.g. compact phases spaces one needs ideas from geometrical quantization, seee.g. [30]. The idea is that (generalized, see below) points of C serve as arguments of the vectorsof the Hilbert space to be constructed.

iii) Preferred Kinematical Poisson SubalgebraConsider the space C∞(C) of smooth functions on C and the space V ∞(C) of smooth vector fieldson C. The vertical polarization of M, that is, the space of fibre coordinates called momentumspace, generates preferred elements of V ∞(C) through (vp[f ])(q) := (p, f)(q) where we havedenoted configuration and momentum coordinates by q, p respectively and v[f ] denotes theaction of a vector field on a function. The pair C∞(C) × V∞(C) forms a Lie algebra definedby [(f, v), (f ′, v′)] = (v[f ′] − v′[f ], [v, v′]) of which the algebra B generated by elements of theform (f, vp) forms a subalgebra. We assume that B is closed under complex conjugation whichbecomes its ∗−operation (involution).

iv) Representation Theory of the Corresponding Abstract ∗−AlgebraWe are looking for all irreducible ∗−representations π : B → L(Hkin) of B as linear operatorson a kinematical Hilbert space Hkin such that the ∗−relations becomes the operator adjointand such that the canonical commutation relations are implemented, that is, for all a, b ∈ B

π(a)† = π(a∗)

[π(a), π(b)] = i~π([a, b]) (I.3.1.1)

Strictly speaking, (I.3.1.1) is to be supplemented by the domains on which the operators aredefined. In order to avoid this one will work with the subalgebra of C∞(C) formed by boundedfunctions, say of compact support and one will deal with exponentiated vector fields in orderto obtain bounded operators. Irreducibility is a physically meaningful requirement becausewe are not interested in Hilbert spaces with superselection sectors and the reason for whywe do not require the full Poisson algebra to be faithfully represented is that this is almostalways impossible in irreducible representations as stated in the famous Groenewald – van Hovetheorem. The Hilbert space that one gets can usually be described in the form L2(C, dµ) whereC is a distributional extension of C and µ is a probability measure thereon. A well-knownexample is the case of free scalar fields on Minkowski space where C is some space of smoothscalar fields on R3 vanishing at spatial infinity while C is the space of tempered distributionson R3 and µ is a normalized Gaussian measure on C.

v) Selection of Suitable Kinematical RepresentationsCertainly we want a representation which supports also the constraints and the Hamiltonian asoperators which usually will limit the number of available representations to a small number,if possible at all. The constraints usually are not in B unless linear in momentum and theexpressions CI := π(CI), H = π(H) will involve factor ordering ambiguities as well as regular-izationand renormalization processes in the case of field theories. In the generic case, CI , Hwill not be bounded and CI will not be symmetric. We will require that H is symmetric andthat the constraints are at least closable, that is, they are densely defined together with their

23

adjoints. It is then usually not too difficult to find a dense domain Dkin ⊂ Hkin on whichall these operators and their adjoints are defined and which they leave invariant. TypicallyDkin will be a space of smooth functions of rapid decrease so that arbitray derivatives andpolynomials of the configuration variables are defined on them and such spaces naturally comewith their own topology which is finer than the subspace topology induced from Hkin whencewe have a topological inclusion Dkin → Hkin.

vi) Imposition of the ConstraintsThe two step process in the classical theory of solving the constraints CI = 0 and looking forthe gauge orbits is replaced by a one step process in the quantum theory, namely looking forsolutions l of the equations CI l = 0. This is because it is obviously solves the constraint at thequantum level (in the corresponding representation on the solution space the constraints arereplaced by the zero operator) and it simultaneously looks for states that are gauge invariantbecause CI is the quantum generator of gauge transformations.

Now, unless the point 0 is in the common point spectrum of all the CI , solutions l to theequations CIl = 0 ∀ I do not lie in Hkin, rather they are distributions. Here one has severaloptions, one could look for solutions in the space D′

kin of continuous linear functionals onDkin (topological dual) or in the space D∗

kin of linear functionals on Dkin with the topologyof pointwise convergence (algebraic dual). Since certainly Hkin ⊂ D′

kin ⊂ D∗kin let us choose

the latter option for the sake of more generality. The topology on Hkin is again finer thanthe subspace topology induced from D∗

kin so that we obtain a Gel’fand triple or Rigged HilbertSpace

Dkin → Hkin → D∗kin (I.3.1.2)

This a slight abuse of terminology since the name is usually reserved for the case that Dkin

carries a nuclear topology (generated by a countable family of seminorms separating the points)and that D∗

kin is its topological dual.

We are now looking for a subspace D∗phys ⊂ D∗

kin such that for its elements l holds

[C ′I l](f) := l(C†

If) = 0 ∀ f ∈ Dkin, ∀I (I.3.1.3)

The prime on the left hand side of this eqution defines a dual, anti-linear representation of theconstraints on D∗

kin. The reason for the adjoint on the right hand side of this equation is thatif l would be an element of Hkin then (I.3.1.3) would be replaced by

[C ′I l](f) :=< CI l, f >kin=< l, C†

If >kin=: l(C†If) ∀ f ∈ Dkin, ∀I (I.3.1.4)

where < ., . >kin denotes the kinematical inner product, so that (I.3.1.3) is the natural extensionof (I.3.1.4) from Hkin to D∗

kin.

vii) AnomaliesSince we have a first class constraint algebra, we know that classically CI , CJ = fIJ

KCK forsome structure functions fIJ

K which depend in general on the phase space point m ∈ M. Thetranslation of this equation into quantum theory is then plagued with ordering ambiguities,because the structure functions turn into operators as well. It may therefore happen that, e.g.

[CI , CJ ] = i~CK fIJK = i~[CK , fIJ K ] + fIJ

KCK (I.3.1.5)

and it follows that any l ∈ D∗phys also solves the equation ([CK , fIJ

K ])′l = 0 for all I, J . If thatcommutator is not itself a constraint again, then it follows that l solves more than only the

24

equations C ′I l = 0 and thus the quantum theory has less physical degrees of freedom than the

classical theory. This situation, called an anomaly, must be avoided by all means.

viii) Dirac Observables and Physical Inner ProductSince generically Hkin∩D∗

phys = ∅, the space D∗phys cannot be equipped with the scalar product

< ., . >kin. It is here wehere Dirac observables come into play. A strong Dirac observableis an operator O on Hkin which is, together with its adjoint, densely defined on Dkin andwhich commutes with all constraints, that is, [O, CI ] = 0 for all I. We require that O is thequantization of a real valued function O on the phase space and the condition just stated is thequantum version of the classical gauge invariance condition O,CI = 0 for all I. A weak Diracobservable is the quantum version of the more general condition O,CI|CJ

= 0∀J = 0 ∀Iand simply means that the space of solutions is left invariant by the natural dual action of theoperator O′D∗

phys ⊂ D∗phys.

A physical inner product on a subset Hphys ⊂ D∗phys is a positive definite sesquilinear form

< ., . >phys with respect to which the O′ become self-adjoint operators, that is, O′ = (O′)⋆

where the adjoint on Hphys is denoted by ⋆. Notice that [O′1, O

′2] = ([O1, O2])

′ so that com-mutation relations on Hkin are automatically transferred to Hphys which then carries a proper∗−representation of the physical observables. The observables themselves will only be definedon a dense domain Dphys ⊂ Hphys and we get a second Gel’fand triple

Dphys → Hphys → D∗phys (I.3.1.6)

In fortunate cases, for instance when the CI are mutually commuting self-adjoint operators onHkin, all we have said is just a fancy way of stating the fact that Hkin has a direct integraldecomposition

Hkin =

∫ ⊕

S

dν(λ)Hλ (I.3.1.7)

over the spectrum S of the constraint algebra with a measure ν and eigenspaces Hλ which areleft invariant by the strong observables and therefore Hphys = H0. In the more general casesthat are of concern to us, more work is required.

ix) Classical LimitIt is by no means granted that the representation Hphys that one finally arrived at, carriessemiclassical states, that is states ψ[m] labelled by gauge equivalence classes [m] of pointsm ∈ M with respect to which the Dirac observables have the correct expectation values andwith respect to which their relative fluctuations are small, that is, roughly speaking

|< ψ[m], O′ψ[m] >phys

O(m)− 1| ≪ 1 and |< ψ[m], (O

′)2ψ[m] >phys

(< ψ[m], O′ψ[m] >phys)2− 1| ≪ 1 (I.3.1.8)

Only when such a phase exists are we sure that we have not constructed some completelyspurious sector of the quantum theory which does not admit the correct classical limit.

I.3.2 Selected Examples with First Class Constraints

In the case that a theory has only first class constraints, Dirac’s algorithm [20] boils down to thefollowing four steps:

25

1)Define the momentum pa conjugate to the configuration variable qa by (Legendre transform)

pa := ∂S/∂qa (I.3.2.1)

where S is the action.2)Equation (I.3.2.1) defines pa as a function of qa, qa and if it is not invertible to define the qa as afunction of qa, pa we get a collection of so-called primary constraints CI , that is, identities among theqa, pa. In this situation one says that S or the Lagrangean is singular.3)Using that qa, pa have canonical Poisson brackets, compute all possible Poisson brackets CIJ :=CI , CJ. If some CI0J0 is not zero when all CK vanish, then add this CI0J0, called a secondaryconstraint, to the set of primary constraints.4)Iterate 3) until the CI are in involution, that is, no new secondary constraints appear.

In this report we will only deal with theories which have no second class constraints, so thisalgorithm is all we need.

Exercise I.3.1.Perform the quantization programme for a couple of simple systems in order to get a feeling for theformalism:

1. Momentum ConstraintM = T ∗R2 with standard Poisson brackets among qa, pa; a = 1, 2 and constraint C := p1.Choose Hkin = L2(R

2, d2x), Dkin = S(R2), D∗kin = S ′(R2) (spaces of functions of rapid decrease

and tempered distributions respectively).Solution:Dirac observables are the conjugate pair q2, p2, Hphys = L2(R, dx2).Hint: Work in the momentum representation and conclude that the general solution is of theform lf(p1, p2) = δ(p1)f(p2) for f ∈ S ′(R).

2. Angular Momentum ConstraintM = T ∗R3 with standard Poisson brackets among qa, pa; a = 1, 2, 3 and constraints Ca :=ǫabcx

bpc. Check the first class property and choose the kinematical spaces as above with R2

replaced by R3.Solution:Dirac observables are the conjugate pair r :=

√δabqaqb ≥ 0, pr = δabq

apb/r, the physical phasespace is T ∗R+ and Hphys = L2(R+, r

2dr) where r is a multiplication operator and pr = i~1rddrr

with dense domain of symmetry given by the square integrable functions f such that f is regularat r = 0.Hint:Introduce polar coordinates and decompose kinematical wave functions into spherical harmonics.Conclude that the physical Hilbert space this time is just the restriction of the kinematical Hilbertspace to the zero angular momentum subspace, that is, Hphys ⊂ Hkin. The reason is of course

that the spectrum of the Ca is pure point (discrete).

3. Relativistic ParticleConsider the Lagrangean L = −m

√−ηµν qµqν where m is a mass parameter, η is the Minkowski

26

metric and µ = 0, 1, .., D. Verify that the Lagrangean is singular, that is, the velocities qµ

cannot be expressed in terms of the momenta pµ = ∂L/∂qµ which gives rise to the mass shellconstraint C = m2 + ηµνpµpν. Verify that this happens because the corresponding action isinvariant under Diff(R), that is, reparameterizations t 7→ ϕ(t), ϕ(t) > 0. Perform the Diracanalysis for constraints and conclude that the system has no Hamiltonian, just the Hamiltonianconstraint C which generates reparameterizations on the kinematical phase space M = T ∗RD+1

with standard Poisson brackets. Now choose kinematical spaces as in 1. with R2 replaced byRD+1.Solution:Conjugate Dirac observables are Qa = qa − q0pa√

m2+δabpapb

and Hphys = L2(RD, dDp) on which

q0 = 0.Hint:Work in the momentum representation and conclude that the general solution to the constraintsis of the form lf = δ(C)f(p0, ~p). Now notice that the δ−distribution can be written as a sum oftwo δ−distribution corresponding to the positive and negative mass shell and choose f to havesupport in the former.

This example has features rather close to those of general relativity.

4. Maxwell TheoryConsider the action for free Maxwell-theory on Minkowski space and perform the Legendretransform. Conclude that there is a first class constraint C = ∂aE

a (Gauss constraint) withLagrange multiplier A0 and a Hamiltonian H =

∫R3 d3x[EaEb+BaBb]/2 where Ea = Aa−∂aA0

is the electric field and Ba = ǫabc∂bAc the magnetic one. Verify that the Gauss constraint gen-erates U(1) gauge transformations A 7→ A − df while Ea is gauge invariant. Choose Hkin tobe the standard Fock space for three massless, free scalar fields Aa and as Dkin, D∗

kin the finitelinear span of n−particle states and its algebraic dual respectively.Solution:Conjugate Dirac observables are the transversal parts of A,E respectively, e.g. Ea

⊥ = Ea −∂a

1∆∂bE

b where ∆ is the Laplacian on R3. The physical Hilbert space is the standard Fock spacefor two free, masless scalar fields corresponding to these transversal degrees of freedom.Hint:Fourier transform the fields and compute the standard annihilation and creation operatorsza(k), z

†a(k) with canonical commutation relations. Express the Gauss constraint operator in

terms of them and conclude that the gauge invariant part satisfies za(k)ka = 0. Introduce

zI(k) = za(k)eaI(k) where ~e1(k), ~e2(k), ~e3(k) := ~k/||k|| form an oriented orthonormal basis.

Conclude that physical states are states without longitudonal excitations and build the Fockspace generated by the z†1(k), z

†2(k) from the kinematical vacuum state.

27

Part II

Mathematical and Physical Foundationsof Quantum General Relativity

28

II.1 Mathematical Foundations

II.1.1 Polarization and Preferred Poisson Algebra BThe first two steps of the quantization programme were already completed in section I.2: The phasespace M is coordinatized by canonically conjugate pairs (Aja, E

aj ) where A is an SU(2) connection

over σ while E is a su(2)−valued vector density of weight one over σ and the Poisson brackets weredisplayed in (I.2.2.6). Strictly speaking, since M is an infinite dimensional space, one must supplyM with a manifold structure modelled on some Banach space but we will skip these functionalanalytic niceties here, see [1] for further information. Also we must specify the principal fibre bundleof which A is the pull-back by local sections of a globally defined connection, and we must specifythe vector bundle associated to that principal bundle under the adjoint representation of which Eis the pull-back by local sections. Again, in order not to dive too deeply into fibre bundle theoreticsubtleties, we will assume that the principal fibre bundle is trivial so that A,E are actually globallydefined. In fact, for the case of G = SU(2) and dim(σ) = 3 one can show that the fibre bundle isnecessarily trivial but for the generalization to the generic case we again refer the reader to [1].

With these remrks out of the way we may begin by defining a polarization. The fact that GRhas been casted into the language of a gauge theory suggests the choice C = A, the space of smoothSU(2) connections over σ.

The next question then is how to choose the space C∞(A). Since we are dealing with a fieldtheory, it is not clear a priori what smooth or even differentiable means. In order to give precisemeaning to this, one really has to equip A with a manifold structure modelled on a Banach space.This is because one usually says that a function F : A → C is differentiable at A0 ∈ A provided thatthere exists a bounded linear functional DFA0 : TA0(A) → C such that F [A0+δA]−F [A0]−DFA0 ·δAvanishes “faster than linearly” for arbitrary tangent vectors δA ∈ TA0(A) at A0. (The proper wayof saying this is using the natural Banach norm on T (A).) Of course, in physicist’s notation thedifferential DFA0 = (δF/δA)(A0) is nothing else than the functional derivative. Using this definitionit is clear that polynomials in Aja(x) are not differentiable because their functional derivative isproportional to a δ−distribution as it is clear from (I.2.2.6). Thus we see that the smooth functionsof A have to involve some kind of smearing of A with test functions, which is generic in field theories.

Now this smearing should be done in a judicious way. The functionF [A] :=

∫σd3xF a

j (x)Aja(x) for some smooth test function F aj of compact support is certainly smooth

in the above sense, its functional derivative being equal to F ja (which is a bounded operator if F

is e.g. an L2 function on σ and the norm on the tangent spaces is an L2 norm). However, thisfunction does not transform nicely under SU(2) gauge transformations which will make it very hardto construct SU(2) invariant functions from them. Here it helps to look up how physicists havedealt with this problem in ordinary canonical quantum Yang-Mills gauge theories and they foundthe following, more or less unique solution [31]:Given a curve c : [0, 1] → σ in σ and a point A ∈ A we define the holonomy or parallel transportA(c) := hc,A(1) ∈ SU(2) as the unique solution to the following ordinary differential equation forfunctions hc,A : [0, 1] → SU(2)

dhc,A(t)

dt= hc,A(t)Aja(c(t))

τj2ca(t), hc,A(0) = 12 (II.1.1.1)

Exercise II.1.1. Verify that (II.1.1.1) is equivalent with

A(c) = P · exp(

c

A) = 12 +∞∑

n=1

∫ t

0

dt1

∫ 1

t1

dt2 ..

∫ 1

tn−1

A(t1)..A(tn) (II.1.1.2)

29

where P denotes the path ordering symbol which orders the curve parameters from left to right ac-cording to their value beginning with the smallest one and A(t) := Aja(c(t))c

a(t)τj/2.

With this definition it is not difficult to verify the following transformation behaviour of A(c)under gauge transformations and spatial diffeomorphisms respectively (recall (I.2.2.9)):

Ag(c) = g(b(c))A(c)g(f(c))−1 and Aϕ(c) = A(ϕ−1(c)) (II.1.1.3)

where b(c), f(c) denote the beginning and final point of a curve respectively. Thus, the behaviourunder gauge transformations is extremely simple which makes it easy to construct gauge invariantfunctions, for instance the Wilson loop functions Tr(A(c)) where c is a closed curve, that is, aloop. This is the reason why QGR is also denoted as Loop Quantum Gravity. That holonomiesalso transform very naturally under spatial diffeomorphisms as depicted in the second equation of(II.1.1.3) has the following mathematical origin: A connection is in particular a one-form, therefore itis naturally integrated (smeared) over one-dimensional submanifolds of σ. Here natural means withoutusing a background metric. Now the holonomy is not really the exponential of

∫cA but almost as

shown in (II.1.1.2). Thus, holonomies are precisely in accordance with our wish to construct abackground independent quantum field theory. Moreover, the simple transformation behaviour underdiffeomorphisms again makes it simple to construct spatially diffeomorphism invariant functions ofholonomies: These will be functions only labelled by diffeomporphism invariance classes of loops, butthese are nothing else than knot classes. Thus QGR has an obvious link with topological quantumfield theory (TQFT) [32] which makes it especially attractive and was one of the major motivationsfor Jacobson, Rovelli and Smolin to consider Wilson loop functions for canonical quantum gravity[33]. Finally one can show [34] that the holonomies separate the points of A, i.e. they encode all theinformation that is contained in a connection.

The fact that the holonomy smears A only one-dimensionally is nice due to the above reasonsbut it is also alarming because its functional derivative is certainly distributional and thus does notexist in the strict mathematical sense. However, in order to obtain a well-defined Poisson algebrait is not necessary to have smooth functions of A, it is only sufficient. The key idea idea is thatif we smear also the electric fields E then we might get a non-distributional Poisson algebra. Byinspection from (I.2.2.6) it is clear that E has to be smeared in at least two dimensions in orderto achieve this. Now again background independence comes to our help: Let ǫabc be the totallyskew, background independent tensor density of weight −1, that is, ǫabc = δ1

[aδ2bδ

3c] where [..] denotes

total antisymmetrization. Then (∗E)jab := Ejab := Ec

jǫabc is a 2-form of density weight 0. ThereforeE is naturally smeared in two dimensions. Notice that the smearing dimensions of momenta andconfiguration variables add up to the dimension of σ, they are dual to each other which is a genericphenomenon for any canonical theory in any dimension. We are therefore led to consider the electricfluxes

Ej(S) =

S

∗Ej (II.1.1.4)

where S is a two-dimensional, open surface. It is easy to check that E(S) := Ej(S)τj has the followingtransformation behaviour

Eg(S) =

S

Adg(∗E) and Eϕ(S) = E(ϕ−1(S)) (II.1.1.5)

Thus, while the transformation under spatial diffeomorphisms is again simple, the one under gaugetransformations is not. However, the idea is that the Ej(S) are the basic building blocks for more

30

complicated functions of E which are already gauge invariant. The prototype of such a function isthe area functional for a parameterized surface XS : D → σ, D ⊂ R2

Ar(S) :=

D

d2u√

det(X∗Sq) (II.1.1.6)

Exercise II.1.2. Define nSa := ǫabcXbS,u1Xc

S,u2 and verify that (II.1.1.6) coincides with

Ar(S) := β

D

d2u√

(Eaj n

Sa )(E

bjn

Sb ) (II.1.1.7)

where β is the Immirzi parameter.

It is clear that Ej(S) =∫Dd2uEa

j nSa so that the area functional can be written as the limit of a

Riemann sum, over small surfaces that partition S, of functions of the electric fluxes for those smallsurfaces.

Let us see whether the Poisson bracket between an electric flux and a holonomy is well-defined.Actually, let us be slightly more general and introduce the following notion: Let us losely think forthe moment of a graph γ as a collection of a finite number of smooth, compactly supported, orientedcurves, called edges e, which intersect at most in their end points, which are called vertices v. Wedenote by E(γ), V (γ) the edge and vertex set of γ respectively. A precise definition will be given insection II.1.2.

Definition II.1.1.Given a graph γ we define

pγ : A → SU(2)|E(γ)|; A 7→ (A(e))e∈E(γ) (II.1.1.8)

A function f : A → C is said to be cylindrical over a graph γ, if there exists a function fγ :SU(2)|E(γ)| → C such that f = fγ pγ. We denote by Cylnγ , n = 0, 1, ..,∞ the set of n−timescontinuously differentiable cylindrical functions over γ and by Cyln the set functions which are cylin-drical over some γ with the same differentiability type. Here we say that f = fγ pγ ∈ Cylnγ if andonly if fγ is n−times continuously differentiable with respect to the standard differential structure onSU(2)|E(γ)|.

Our Poisson Algebra will be based on the set of functions Cyl∞ which certainly form an AbeleanPoisson subalgebra. Our next task will be to compute the Poisson bracket between a flux and anelement of Cyl∞. In order to compute this we will use the chain rule (f ∈ Cyl∞γ )

Ej(S), f(A) =∑

e∈E(γ)

[∂fγ(he′e′∈E(γ))

∂(he)ABEj(S), (he)AB]|he′=A(e′) (II.1.1.9)

so that the bracket will be well-defined once the bracket between a holonomy and a flux is well-defined. To compute this the intersection structure of e with S is somewhat important. In order tosimplify the notation, we notice that we can always take γ to be adapted to S, that is, every edge ebelongs to one of the following three types:a) e ∈ Eout(γ) ⇔ e ∩ S = ∅.b) e ∈ Ein(γ) ⇔ e ∩ S = e.c) e ∈ Etrans(γ) ⇔ e ∩ S = b(e).This can be achieved by subdividing edges into a finite number of segments and inverting their

31

S

n^S

e_in

e_out

e_trans, σ=−1

e_trans, σ=+1

Figure 7: Intersection structure of surfaces with edges

orientation if necessary as depicted in figure 7 (strictly speaking, this is true only if S is compactlysupported, open, oriented and analytic). We also need to introduce the function σ(S, e) whichvanishes for e ∈ Ein(γ) ∪ Eout(γ) and which is ±1 for e ∈ Etrans(γ) if the orientations of S and eagree or disagree respectively. The easiest case is e ∈ Etrans(γ), σ(S, e) = 1. We find

Ej(S), A(e) = κ

D

d2u nSa (u)

∫ 1

0

dsea(s)δ(XS(u), e(s))A(es)τj2A(es)

−1A(e) (II.1.1.10)

where es(t) := e(st). Noticing that the support of the δ−distribution is at XS(u) = e(0) which is aninterior point of S but a boundary point of e, a careful analysis reveals that (II.1.1.10) reduces to

Ej(S), A(e) =κ

4τjA(e) (II.1.1.11)

With this result, (II.1.1.9) can be written in the compact form

Ej(S), f(A) =κ

4

e∈E(γ)

σ(S, e)[Rjefγ]|he′=A(e′) (II.1.1.12)

where we have defined the right invariant vector fields

(Rjefγ)(he′e′∈E(γ)) := (

d

dt)t=0fγ(he′e′ 6=e, etτjhe) (II.1.1.13)

We can now define the vector fields vjS on Cyl∞ by vjS[f ] := Ej(S), f and arrive at the Poisson∗−algebra B generated by the vjS, f ∈ Cyl∞ with involution defined by complex conjugation throughthe general formula [(f, v), (f ′, v′)] = (v[f ′] − v′[f ], [v, v′]).

Exercise II.1.3.Fill the gaps between (II.1.1.9) and (II.1.1.12.Hint:Use formula II.1.1.2 in order to derive (II.1.1.10), then expand XS(u) − e(t) around u = u0 definedby XS(u0) = e(0) and t = 0 to linear order in u − u0 and sufficiently high order in t to arrive at(II.1.1.11). (Notice that e is only transversal, so e(0) may be tangential to S in e(0) !) Verify thatthe end result coincides with (II.1.1.12).

32

So we see that we arrive at a well defined algebra B by smearing the momenta in two dimensions.Could we also smear them in three dimensions ? The answer is negative: Consider a one-parameterfamily of surfaces t 7→ St and define Ej(S) :=

∫dt Ej(St). Then f 7→ Ej(S), f maps f out

of Cyl∞ because it involves an integral over t and thus depends on an uncountably infinite numberof edges rather than a finite number. Thus this algebra would not be closed so that if we wouldlike to stick with at least countably infinite graphs then we are forced to stick with two dimensionalsmearings of the electric fluxes !

II.1.2 Representation Theory of B and Suitable Kinematical Represen-tations

The representation Theory of B has been considered only rather recently [35] and the analysis is notyet complete. However, if one sticks to irreducible representations for which 1) the flux operators arewell-defined and self-adjoint (in other words, the corresponding one parameter unitary groups areweakly continuous) and 2) the representation is spatially diffeomorphism invariant, then the uniquesolution to the representation problem is the representation which we describe in this section.

This representation is of the form H0 = L2(A, dµ0) where A is a certain distributional extensionof A and µ0 is a probability measure thereon. The most elegant description of this Hilbert spaceuses the theory of C∗−algebras [36] but fortunately there is a purely geometric description available[37] which is easier to access for the beginner. In what follows we assume for simplicity that σ is anoriented, connected, simply connected smooth manifold. One can show that each smooth manifoldadmits at least one analytic structure (i.e. the atlas of charts consists of real analytic maps) and weassume to have picked one once and for all.

II.1.2.1 Curves, Paths, Graphs and Groupoids

Definition II.1.2.i)By a curve c we mean a map c : [0, 1] → σ which is piecewise analytic, continuous, oriented and anembedding (does not come arbitrarily close to itself). It is automatically compactly supported. Theset of curves is denoted C in what follows.ii)On C we define maps , (.)−1 called composition and inversion respectively by

[c1 c2](t) =

c1(2t) t ∈ [0, 1

2]

c2(2t− 1) t ∈ [12, 1]

(II.1.2.1)

if f(c1) = b(c2) andc−1(t) = c(1 − 2t) (II.1.2.2)

iii)By a path p we mean an equivalence class of curves c which differ from each other by a finite number ofreparameterizations and retracings, that is, c ∼ c′ if there either exists a map t 7→ f(t), f(t) > 0 withc = c′f or we may write c, c′ as compositions of segments in the form c = s1s2, c

′ = s1s3s−13 s2

(and finite combinations of such moves). Notice that a curve induces its orientation and its end pointson its corresponding path. The set of paths is denoted by P.iv)By a graph γ we mean a finite collection of elements of P. We may break paths into pieces such that

33

l

e’

e

e e^−1

e_1

v_1

v_2

v_M

e_N

e_2

Figure 8: Paths and graphs

γ can be thought of as a collection of edges e ∈ E(γ), that is, paths which have an entire analyticrepresentative and which intersect at most in their end points v ∈ V (γ) called vertices. The set ofgraphs is denoted by Γ.

These objects are depicted in figure 8.

Exercise II.1.4.a)Despite the name, composition and inversion does not equip C with a group structure for manyreasons. Verify that composition is not associative and that the curve c c−1 is not simply b(c) butrather a retracing. Moreover, contemplate that C does not have a unit and that not every two elementscan be composed.b)Define composition and inversion of paths by taking the equivalence class of the compositions andinversions of any of their representatives and check that this definition is well defined. Check thatthen composition of paths is associtive and that p p−1 = b(p). However, P still does not have a unitand still not every two elements can be composed.c)Let Obj:= σ and for each x, y ∈ Σ let Mor(x, y) := p ∈ P : b(p) = x, f(p) = y. Recall themathematical definition of a category and conclude that P is a category in which every morphism isinvertible, that is, a groupoid.d)Define the relation ≺ on Γ by saying that γ ≺ γ′ if and only if every e ∈ E(γ) is a finite compositionof the e′ ∈ E(γ′) and their inverses. Verify that ≺ equips Γ with the structure of a directed set, thatis, for each γ, γ′ ∈ Γ we find γ′′ ∈ Γ such that γ, γ′ ≺ γ′′.Hint:For this to work, analyticity of the curve representatives is crucial. Smooth curves can intersect inCantor sets and thus define graphs which are no longer finitely generated. Show first that this is notpossible for analytic curves.e)Given a curve c with path equivalence class p notice that for the holonomy with respect to A ∈ A holdsA(c) = A(p). Contemplate that, in particular, every group is a groupoid and that every connectionA ∈ A qualifies as a groupoid homomorphism, that is, A : P → SU(2); p 7→ A(p) with

A(p p′) = A(p)A(p′) and A(p−1) = (A(p))−1 (II.1.2.3)

The fact that holonomies are really defined on paths rather than curves and that holonomies arecharacterized algebraically by II.1.2.3 makes the following definition rather natural.

34

Definition II.1.3. The quantum configuration space is defined as the set A := Hom(P, SU(2)) ofall algebraic, arbitrarily non-continuous groupoid morphisms.

Here non-continuous means that in contrast to A ∈ A for an element A ∈ A it is possible thatA(p) = 1 varies discontinuously as we vary p continuously. Thus, A can be thought of as a distribu-tional extension of A.

II.1.2.2 Topology on ASo far A is just a set. We now equip it with a topology. The idea is actually quite simple. Recall themaps (II.1.1.8) which easily extend from A to A and maps A into SU(2)|E(γ)|. Now SU(2)|E(γ)| is acompact Hausdorff topological group5 in its natural manifold topology and we would like to exploitthat. Thus we are motivated to consider the spaces Xγ := Hom(γ, SU(2)|E(γ)|) where γ is consideredas a subgroupoid of Γ with objects V (γ) and morphisms generated by the e ∈ E(γ). The map

Xγ → SU(2)|E(γ)|; xγ 7→ xγ(e)e∈E(γ) (II.1.2.4)

identifies Xγ with G|E(γ)| since xγ ∈ Xγ is already defined by which values it takes on the e ∈ E(γ)and we may thus use this identification in order to equip Xγ with a compact Hausdorff topology.Now consider the uncountably infinite product

X∞ :=∏

γ∈Γ

Xγ (II.1.2.5)

A standard result from general topology, Tychonov’s theorem, tells us that the smallest topology onX∞ such that all the maps pγ : X∞ → Xγ; (xγ)γ∈Γ 7→ xγ are continuous is a compact Hausdorfftopology6. Now we would like to identify A with X∞ through the restriction map

Φ′ : A → X∞; A 7→ (xγ := A|γ = pγ(A))γ∈Γ (II.1.2.6)

However, that map cannot be surjective because the points of A satisfy the following constraintwhich encodes the algebraic properties of a generalized connection: Let γ ≺ γ′ and define the graphrestriction maps

pγ′γ : Xγ′ → Xγ ; xγ′ 7→ (xγ′)|γ (II.1.2.7)

which satisfy the compatibility condition

pγ′′γ = pγ′γ pγ′′γ′ for γ ≺ γ′ ≺ γ′′ (II.1.2.8)

Then automaticallypγ′γ(A|γ′) = A|γ (II.1.2.9)

We are therefore forced to consider the subset of X∞ defined by

X := (xγ)γ∈Γ ∈ X∞; pγ′γ(x′γ) = xγ ∀ γ ≺ γ′ (II.1.2.10)

5Here it is crucial that G = SU(2) is compact and thus for non-real Immirzi parameter all of what follows wouldbe false.

6Recall that we know the topology on a space when we know a base of open sets from which we obtain all opensets by arbitrary unions and finite intersections. Since the preimages of open sets under continuous functions are openby definition, we obtain a topology once we know which functions are continuous.

35

Exercise II.1.5.i)Show that the maps (II.1.2.7) are continuous surjections.Hint:Exploit the identification of the Xγ with powers of SU(2) and the continuity of multiplication andinversion in groups to establish continuity. To establish surjectivity use the fact that each edge e ofγ contains an edge e′e of γ′ as a segment such that the e′e do not overlap each other. Now given xγset xγ′(e

′e) = xγ(e) and extend trivially away from the e′e. Check that this defines an element of Xγ′.

ii)Show that X is a closed subset of X∞.Hint:Since X is not a metric space we must work with nets and show that every net of points xα ∈ Xwhich converges in X∞ actually converges in X. Using the defintion of the topology on X∞, showthat this is equivalent to showing that the pγ(x

α) = xαγ converge to points xγ which satisfy (II.1.2.9)and verify this using continuity of the pγ′γ just established.

The surjectivity of the pγ′γ qualifies X as the so-called projective limit of the Xγ, a mathematicalstructure which is independent of our concrete context once we have a directed index set Γ at ourdisposal and surjective projections which satisfy the compatibility condition (II.1.2.8).

Now another standard result from topology now tells us that X, being the closed subset of acompact Hausdorff space, is a compact Hausdorff space in the subspace topology and the questionarises whether

Φ : A → X; A 7→ (xγ := A|γ = pγ(A))γ∈Γ (II.1.2.11)

is a bijection. Injectivity is fairly easy to see while surjectivity is a little bit tricky.

Exercise II.1.6.Show that (II.1.2.11) is a bijection.Hint:Given x ∈ X and p ∈ P choose any γp ∈ Γ such that p ∈ γp and define Ax by Ax(p) := xγp(p). Showthat this definition is well defined using the directedness of Γ and that Ax is a groupoid homomorphism.

Let us collect these results in the following theorem [38].

Theorem II.1.1.The space A equipped with the weakest topology such that the maps pγ of (II.1.1.8) are continuous,is a compact Hausdorff space.

The value of this result is that it gives us a powerful tool for constructing measures on A.

II.1.2.3 Measures on AA powerful theorem due to Riesz and Markov, sometimews called the Riesz representation theorem,tells us that there is a one – to – one correspondence between the positive linear functionals Λon the algebra C(A) of continuous functions on a compact Hausdorff space A and (regular, Borel)probability measures µ thereon through the simple formula

Λ(f) :=

Adµ(A) f(A) (II.1.2.12)

36

One says Λ is represented by f . Here a linear functional is called positive if Λ(|f |2) ≥ 0 forany f ∈ C(A). A function algebra on a compact space can be equipped with the sup – norm||f || := supA∈A |f(A)| which evidently has the so-called C∗−property ||ff || = ||f ||2 so that (w.l.g.we may take C(A) to be complete w.r.t. the norm) C(A) is a C∗−algebra. A standard result in func-tional analysis reveals that positive linear functionals on C∗−algebras are automatically continuous,|Λ(f)| ≤ Λ(1) ||f || and if we choose the normalization of Λ to be Λ(1) = 1 then µ is a probabilitymeasure.

In order to specify the measure µ0 that we are interested in, it is therefore enough to specify apositive linear functional Λ0. The most elegant way of defining Λ0 is through the following definition.

Definition II.1.4.i)Given a graph γ, label each edge e ∈ E(γ) with a triple of numbers (je, me, ne) where je ∈ 1

2, 1, 3

2, 2, ..

is a half-integral spin quantum number and me, ne ∈ −je,−je + 1, .., je are magnetic quantumnumbers. A quadruple

s := (γ,~j := jee∈E(γ), ~m := mee∈E(γ), ~n := nee∈E(γ)) (II.1.2.13)

is called a spin network (SNW). We also write γ(s) etc. for the entries of a SNW.ii)Choose once and for all one representative ρj , j > 0 half integral, from each equivalence class ofirreducible representations of SU(2). Then

Ts : A → C; A 7→∏

e∈E(γ)

[√

2je + 1[ρje(A(e))]mene] (II.1.2.14)

is called the spin-network function (SNWF) of s. Here [ρj(.)]mn denotes the matrix elements of thematrix valued function ρj(.).

An example of a SNW, which can be arbitrarily large and with vertices of arbitrarily high valence,is given in figure 9. The original motivation for the definition of spin network functions [40] in loopquantum gravity was the fact that they are linearly independent in contrast to the Wilson loopfunctions which suffer from the so-called Mandelstam identities. For SU(2) matrices h, h′ they areTr(h) Tr(h′) = Tr(hh′)+Tr(h(h′)−1) and Tr(h) = Tr(h−1) which leads to an infinite tower of identitiesof the form

[Tr(A(α1)) Tr(A(α2))]Tr(A(α1)) = Tr(A(α1))[Tr(A(α2)) Tr(A(α1))] (II.1.2.15)

depending on how we bracket the product of traces involving the three loops α1, α2, α3 with a commonbase point. The SNWF’s remove these cumbersome identities first by labelling functions by edgesrather than loops and secondly by the simple observation that a tensor product of (fundamental)representations can be uniquely decomposed into irreducibles (Clebsh-Gordon decomposition).

Theorem II.1.2.The uniform (Ashtekar – Lewandowski) measure µ0 is uniquely defined by the positive linear func-tional [39]

Λ0(Ts) :=

1 s = (∅,~0,~0,~0)0 otherwise

(II.1.2.16)

37

j_1

j_2

j_54

j_3

j_4

j_5

j_6

j_7

j_8

j_9

j_10

j_11

j_12

j_13j_14

j_15

j_16

j_17

j_18

j_19

j_20

j_21

j_22

j_23

j_24

j_25

j_26

j_27

j_28

j_29

j_30

j_31

j_32

j_33

j_34

j_35

j_36

j_37

j_38

j_39

j_40

j_41

j_42

j_43

j_44

j_45

j_46j_47

j_48

j_49

j_50

j_51

j_52

j_0

j_53

Figure 9: A SNW. Orientations and magnetic quantum numbers are suppressed

Exercise II.1.7.i)Recall the representation theory of SU(2) from the quantum mechanics of angular momentum andverify that the SNWF are indeed linearly independent.ii)Verify that Λ0 is a positive linear functional.Hint:Using the Stone – Weierstrass theorem, show first that the finite linear combinations of SNWF aredense in C(A). By continuity of Λ0 it is therefore sufficient to check positivity on finite linearcombinations

f =N∑

n=1

znTsn, N <∞, zn ∈ C (II.1.2.17)

with sn mutually different SNW’s. To see this, verify that Λ0(TsTs′) = 0 for s 6= s′ by using theClebsh – Gordon formula j ⊗ j′ ≡ (j + j′) ⊕ (j + j′ − 1) ⊕ ..⊕ (|j − j′|).iii)For the representation theory of compact groups fundamental is a theorem due to Peter and Weyl[41] which for SU(2) amounts to saying that the functions

Tjmn : SU(2) → C; h 7→√

2j + 1[ρj(h)]mn (II.1.2.18)

form an orthonormal basis for the Hilbert space L2(SU(2), dµH) where µH is the normalized Haarmeasure on SU(2) (the unique normalized measure which invariant under inversion as well as leftand right translation in SU(2)). Based on this result, show that the SNWF form an orthonormalbasis for the Hilbert space L2(A, dµ0).

Let us summarize the results of the exercise in the following theorem [40].

Theorem II.1.3.The kinematical Hilbert space Hkin := L2(A, dµ0) defined by (II.1.2.16) is non-separable and has theSNWF’s Ts as orthonormal basis.

38

II.1.2.4 Representation Property

So far we did not verify that Hkin is a representation space for our ∗−algebra B of basic operators.This will be done in the present section. Indeed, until today no other irreducible representation ofthe holonomy – flux algebra has been found (except if one allows also infinite graphs [42]).

By theorem (II.1.4) the subspace of finite linear combinations of SNWF’s is dense in Hkin withrespect to the L2 norm. On the other hand, we notice that the definition of Cyl∞(A) simply extendsto Cyl∞(A) and that finite linear combinations of SNWF’s form a subspace of Cyl∞(A). Thus, wemay choose Dkin := Cyl∞(A) and obtain a dense, invariant domain of B as we will see shortly. Wedefine a representation of the holonomy – flux algebra by (f ′ ∈ Cyl∞(A), f ∈ Cyl∞(A), A ∈ A)

[π(f) · f ′](A) := (f ′f)(A)

[π(vjS) · f ](A) = [π(vjS)π(f) · 1](A) = [([π(vjS), π(f)] + π(f)π(Ej(S))) · 1](A)

:= i~[π(vjS[f ]) · 1](A) = i~(vjS[f ])(A) (II.1.2.19)

Thus π(f) is a multiplication operator while π(vjS) is a true derivative operator, i.e. it annihilatesconstants. Notice that the canonical commutation relations are already obeyed by construction, thuswe only need to verify the ∗−relations and the fact that π(vjS) annihilates constants will be crucialfor that.

The π(f) are bounded multiplication operators (recall that smooth, i.e. in particular continuous,functions on copmpact spaces are uniformly bounded, that is, have a sup – norm) so that the adjointis complex conjugation, therefore there is nothing to check. As for π(vjS) we notice that given twosmooth cylindrical functions on A we always find a graph γ over which both of them are cylindricaland which is already adapted to S.

Exercise II.1.8.Let f be cylindrical over γ. Verify that

Λ0(f) =

SU(2)|E(γ)|

e∈E(γ)

dµH(he)fγ(hee∈E(γ)) (II.1.2.20)

Hint:Write f as a (Cauchy limit of) finite linear combinations of SNWF’s and verify that (II.1.2.20)coincides with (II.1.2.16).

Using the explicit expression (II.1.1.12) and the result of exercise II.1.8 it is easy to see that thesymmetry condition < f, π(vjS)f

′ >kin=< π(vjS)f, f′ >kin is equivalent with the condition

< F,RjF ′ >L2(SU(2),dµH )= − < RjF, F ′ >L2(SU(2),dµH ) (II.1.2.21)

for any F, F ′ ∈ C∞(SU(2)) and Rj is the right invariant vector field on SU(2). However, µH is bydefinition invariant under left translations and Rj is a generator of left translations in SU(2) so theresult follows. This shows that Dkin is contained in the domain of π(vjS)

† and that the restriction ofthe adjoint to Dkin coincides with π(vjS). That Dkin is actually a domain of (essential) self-adjointnessrequires a little bit more work but is not difficult to see, e.g. [1].

Finally, let us verify that the representation is irreducible. By definition, a representation isirreducible if every vector is cyclic and a vector Ω is cyclic if the set of vectors π(a)Ω, a ∈ B is dense.Now the vector Ω = 1 is cyclic because the vectors π(f)Ω = f, f ∈ Cyl∞ are already dense. Givenan arbitrary element ψ ∈ Hkin we know that it is a Cauchy limit of finite linear combinations of spin

39

network functions. Thus, if we can show that we find a sequence an ∈ B such that π(an)ψ convergesto Ω, then we are done. It is easy to see (exercise) that this problem is equivalent to showing thatany F ∈ L2(F, dµH) can be mapped by the algebra formed out of right invariant vector fields andsmooth functions on SU(2) to the constant function.

Exercise II.1.9.Check that this is indeed the case.Hint:Show first that it is sufficient to establish that any polynomial p of the a, b, c, d, ad − bc = 1 for

h =

(a bc d

)∈ SU(2) can be mapped to the constant function. Show then that suitable linear

combinations of the Rj , j = 1, 2, 3 with coefficients in C∞(SU(2)) produce the derivatives ∂a, ∂b, ∂cand convince yourself that aNp is a polynomial in a, b, c for sufficiently large N .

Let us collect these results in the following theorem [43].

Theorem II.1.4.The relations (II.1.2.19) define an irreducible representation of B on Hkin.

Thus, the representation space Hkin constructed satisfies all the requirements that qualify it asa good kinematical starting point for solving the quantum constraints. Moreover, the measure µ0

is spatially diffeomorphism invariant as we will see shortly and together with the uniqueness resultquoted at the beginning of this section, this is the only representation with that property. Thereare, however, doubts on physical grounds whether one should insist on a spatially diffeomorphisminvariant representation because the smooth and even analytic structure of σ which is encoded inthe spatial diffeomorphism group should not play a fundamental role at short scales if Planck scalephysics is fundamentally discrete. In fact, as we will see later, QGR predicts a discrete Planckscale structure and therefore the fact that we started with analytic data and ended up with discrete(discontinuous) spectra of operators looks awkward. Therefore, on the one hand we should keep inmind that other representations are possibly better suited in the final picture, on the other handthere is no logical contradiction within the present formulation and in fact in 2+1 gravity one has afinal combinatorical description while one started with analytical structures as well.

II.2 Quantum Kinematics

In this section we discuss the complete solution of the Gauss and Vector constraint as well as thequantization of kinematical, geometrical operators that measure the length, area and volume ofcoordinate curves, surfaces and regions respectively. We call these results kinematical because theGauss and Vector constraint do not generate dynamics, this is the role of the Hamiltonian constraintwhich we will discuss in the third part. Moreover, the kinematical geometrical operators do notcommute with the Vector constraint or the Hamiltonian constraint and are therefore not Diracobservables. However, as we will show, one can turn these operators easily into Dirac observables,at least with respect to the Vector constraint, and the fact that the spectrum is discrete is robustunder those changes.

40

e_1

e_3

e_2

e_1e_2

e_3

j_1,m_1,n_1

j_1,m_1,n_1

ϕ

Figure 10: Action of Spatial Diffeomorphisms on SNW’s

II.2.1 The Space of Solutions to the Gauss and Spatial Diffeomorphism

Constraint

Recall the transformation behaviour of classical connections A ∈ A under SU(2) gauge transforma-tions and spatial diffeomorphisms (II.1.1.3). These equations trivially lift from A to A and we mayconstruct corresponding operators as follows: Let G := Fun(Σ, SU(2)) be the set of local gauge trans-formations without continuity requirement and consider the set Diffω(Σ) of analytic diffeomorphisms.We are forced to consider analytic diffeomorphisms as otherwise we would destroy the analyticityof the elements of Γ. These two groups have a natural semi – direct product structure that has itsorigin in the algebra (I.2.2.7) and is given by

[G Diffω(Σ)] × [G Diffω(Σ)] → [G Diffω(Σ)]; [g, ϕ] · [g′, ϕ′] = [(g′ ϕ−1)g, ϕ ϕ′] (II.2.1.1)

Exercise II.2.1.Verify (II.2.1.1).Hint:Define [g, id] · A := Ag, [id, ϕ] · A := Aϕ and [g, ϕ] · A := [g, id] · ([id, ϕ] ·A).

We now define representations

U : G → B(Hkin); g 7→ U(g)

V : Diffω(Σ) → B(Hkin); ϕ 7→ V (ϕ) (II.2.1.2)

densely on f = p∗γfγ ∈ Dkin by

[U(g)f ](A) := fγ(g(b(e)) A(e) g(f(e))−1e∈E(γ))[V (ϕ)f ](A) := fγ(A(ϕ−1(e))e∈E(γ)) (II.2.1.3)

Here B(.) denotes the bounded operators on a Hilbert space. This definition of course comes preciselyfrom the classical formula (II.1.1.3). The action of a diffeomorphism on a SNWF Ts is thereforesimply by mapping the graph γ(s) to ϕ−1(s) while the labels je, me, ne are carried from e to ϕ−1(e)as depicted in figure 10. Then the following theorem holds [43].

Theorem II.2.1.The relations (II.2.1.3) define a unitary representation of the semi – direct product kinematical groupG × Diffω(Σ).

Exercise II.2.2.Prove theorem (II.2.1).Hint:Check unitarity on the SNWF basis using the bi – invariance of the Haar measure. That (II.2.1.3)holds can be traced back to exercise II.2.1.

41

The unitarity of the kinematical gauge group implies invariance of the measure µ0 and thussupplies additional motivation for the representation space Hkin. Notice that the statement that(II.2.1.3) defines a representation in particular means that the kinematical constraint algebra is freeof anomalies. This should be contrasted with string theory where the anomly sits also in the spatialdiffeomorphism group (e.g. Diff(S1) for the closed string) unless one chooses the critical dimensionD = 25(9) for the bosonic (supersymmetric) string.

Let us now solve the kinematical constraints. By definition, we are supposed to find algebraicdistributions l ∈ D∗

kin which satisfy

l(U(g)f) = l(V (ϕ)f) = l(f) ∀ g ∈ G, ϕ ∈ Diffω(Σ), f ∈ Dkin (II.2.1.4)

Now it is not difficult to see that any element of D∗kin can be conveniently written in the form

l(.) =∑

s

cs < Ts, . >kin (II.2.1.5)

where cs are complex valued coefficients and the uncountably infinite sum extends over all possibleSNW’s. The general solution to (II.2.1.4) is then easy to describe: Invariance under G means thatfor fixed γ the coefficients cγ,~j,~m,~n have to be chosen, as ~j, ~m,~n vary, in such a way that at each vertexof γ the resulting function is gauge invariant. That is, if j1, .., jn are the labels of edges incident at v,then the cs have to arrange themselves to a projector on the trivial representations contained in thetensor product j1 ⊗ ..⊗ jn. Such a projector is also called intertwiner in the mathematical literature.For SU(2) this leads to the theory of Clebsh-Gordon coefficients, 6j−symbols etc. As for Diffω(Σ)we see that cϕ(γ),~j,~m,~n must be independent of ϕ, therefore cϕ,~j,~m,~n depends only on the generalizedknot class of γ ! We say generalized because, as we will see later on, the physically relevant graphsare those with self-intersections while classical knot theory deals only with smooth curves.

One may ask whether one should already define a physical inner product with respect to theGauss and spatial Diffeomorphism constraint and then solve the Hamiltonian constraint in a second,separate step on that already partly physical Hilbert space . While such a Hilbert space can indeedbe constructed [43] it is of no use for QGR because the Hamiltonian constraint cannot leave thatHilbert space invariant as we see from the second equation in (I.2.1.12) and we must construct thephysical inner product from the full solution space to all constraints. However, at least with respectto the kinematical constraints the full quantization programme including the question of observableshas already been completed except for the analysis of the classical limit.

II.2.2 Kinematical Geometrical Operators

We will restrict ourselves to the description of the area operator the classical expression of which wealready wrote in (II.1.1.6) and (II.1.1.7).

In order to quantize Ar(S) one starts from (II.1.1.7) and decomposes the analytical, compactlysupported and oriented surface S or, equivalently, its preimage D under XS into small pieces SI .Then the exact area functional is approximated by the Riemann sum

Ar(S) = β∑

I

√Ej(SI)2 (II.2.2.1)

This function is easily quantized because Ej(SI) = i~vjs is a self-adjoint operator so that the sumover j of its squares is positive semi-definite, hence its square root is well-defined. Let us denote the

42

resulting, partition dependent operator by Ar(S). Now one can show that the (strong) limit as

the partition is sent to the continuum exists [44] and a partition independent operator Ar(S) results[44].

Theorem II.2.2.The area functional admits a well-defined quantization Ar(S) on Hkin with the following properties:

i) Ar(S) is positive semidefinite, (essentially) self-adjoint with Cyl2(A) as domain of (essential)self-adjointness.

ii) The spectrum Spec(Ar(S)) is pure point (discrete) with eigenvectors being given by finite linearcombinations of spin network functions.

iii) The eigenvalues are given explicitly by

λj1,j2,j12 =βℓ2P4

√2j1(j1 + 1) + 2j2(j2 + 1) − j12(j12 + 1

J12 ∈ j1 + j2, j1 + j2 − 1, .., |j1 − j2| (II.2.2.2)

where j1, j2 are spin quantum numbers and ℓ2P = ~κ is the Planck area. The spectrum has anarea gap (smallest non-vanishing eigenvalue) given by

λ0 = βℓ2P

√3

4(II.2.2.3)

iv) Spec(Ar(S)) contains information about the topology of S, for instance it matters whether∂S = ∅ or not.

Exercise II.2.3.Verify that the area gap is indeed given by (II.2.2.3) and check that the distance between subsequenteigenvalues rapidly decreases as j1, j2 → ∞. Can one give an asymptotic formula for N(A), thenumber of eigenvalues (discarding multiplicity) in the interval [A− ℓ2P , A+ ℓ2P ] ? Thus, a correspon-dence principle, important for the classical limit is valid. If the spectrum would only consist of the

main series λj =ℓ2P2

√j(j + 1) which one obtains for j1 = j2 = j, j12 = 0 then such a correspondence

principle would certainly not hold which is, e.g., relevant for the black body spectrum of the Hawkingradiation.

Theorem II.2.2 is an amazing result for several reasons:A)First of all, the expression for Ar(S) depends non-polynomially, not even analytically on the productEaj (x)E

bj (x), x ∈ S. Now Ea

j (x) becomes an operator valued distribution in the quantum theory

and products of distributions at the same point are usually badly divergent. However, Ar(S) isperfectly well-defined ! This is the first pay-off for sticking to a rigorous and background independentformalism !B)

Although S, γ,Σ, .. are analytical, the spectrum Spec(Ar(S)) is discrete. In other words, suppose weare measuring the area of a sheet of paper with a spin-network state. As long as the sheet does notcut an edge of the graph, the area eigenvalue is exactly zero no matter how “close” the edge andthe sheet are. We have put the word “close” in inverted commas because this word has no meaning:

43

Since there is no background metric, we do not know what close means, only diffeomorphism invariantnotions have a meaning such as “the edge is cut” or “the edge is not cut”. However, once the edgeis cut the area eigenvalue jumps at least by the area gap. This strongly hints that the microscopicalgeometry is really distributional (discontinuous) and that we have a discrete Planck scale structure,the role of the atoms of geometry being played by the one-dimensional (polymer-like) excitationslabelled by SNW’s. One may speculate that this discrete structure is fundamental and that theanalyticity assumptions that we began with should be unimportant, in the final picture everythingshould be only combinatorical. The smooth geometry that we are familiar with at macroscopic scalesis merely a result of coarse graining, for instance in order that a SNWF labelled with spin j = 1/2on every edge assigns to our sheet of paper its area of about 100cm2, an order of 1068 edges of theSNW have to cut the sheet !C)

Qualitatively similar results apply to the volume operator Vol(R) [44, 45] and the length operator

Len(c) [46] whose classical expressions are given by

Vol(R) =

R

d3x√

det(q) and Len(c) =

c

√qabdxadxb (II.2.2.4)

D)These kinematical operators are certainly not Dirac observables because they are not even spatiallydiffeomorphism invariant (but SU(2) invariant) since the objects R, S, c are just coordinate subman-ifolds of Σ. Thus, one may wonder whether the properties of the spectrum just stated have anysignificance at all. The answer is believed to be affirmative as the following argument shows: Forinstance, instead of Vol(R) consider

VolEM =

Σ

d3x√

det(q) θ(qab√det(q)

[EaEb +BaBb]) (II.2.2.5)

where we have coupled a Maxwell field to GR with electromagnetic fields Ea, Eb and θ is the stepfunction. The physical meaning of (II.2.2.5) is that it measures the volume of the region wherethe electromagnetic field energy density is non-vanishing and it is easy to check that (II.2.2.5) isactually spatially diffeomorphism invariant ! Now in QGR the argument of the step function canbe given a meaning as an operator (valued distribution) as we will see in the next section and thetheta function of an operator can be defined through the spectral theorem. Since the spectrum ofthe θ−function consists only of 0, 1, the spectrum of (II.2.2.5) should actually coincide with that

of Vol(R) [47]. A similar argument should also be valid with respect to Dirac observables commutingwith the Hamiltonian constraint.E)The existence of the area gap is also at the heart of the finiteness of the Bekenstein – Hawkingentropy of black holes as we will see.

44

Part III

Selected Areas of Current Research

45

III.1 Quantum Dynamics

The Hamiltonian constraint C of QGR is, arguably, the holy grail of this approach to quantum gravity,therefore we will devote a substantial amount of space to this subject. In fact, unless one can quantizethe Hamiltonian constraint, literally no further progress can be made so that it is important to knowwhat its status is. From the explicit, non-polynomial expression (I.2.2.5) it is clear that a well-defined operator version of this object will be extremely hard to obtain and in fact this had beenthe major obstacle in the whole approach until the mid 90’s. In particular, within the original ADMformulation only formal results were available. However, since with the new connection formulationalso the non-polynomial kinematical operators of the previous section could be constructed, chancesmight be better.

At this point we include a brief account of the historical development of the subject in order toavoid confusion as one looks at older papers:Originally one chose the Immirzi parameter as β = ±i and considered C =

√det(q) rather than C

because then C is actually a simple polynomial of only fourth order (the “More” term disappears).Polynomiality was considered as mandatory. There were three problems with this idea:1) The non-polynomiality was shifted from C into the reality conditions A + A = 2Γ(E) wherethe spin connection Γ is now a highly non-polynomial function of E. The operator version of thisequation should be very hard to implement.2) If A is complex, then we are dealing with an SL(2,C) bundle rather than an SU(2) bundle. SinceSL(2,C) is not compact, the mathematical apparatus of section II is blown away.3) Even formal trials to quantize C resulted in either divergent, or background dependent operators.In [27] it was suggested to keep β real which solves problems 1) and 2), however, then C becomeseven more complicated and anyway problem 3) is not cured. Finally in [48] it was shown that non-polynomiality is not necessarily an obstacle, even better, it is actually required in order to arrive at awell-defined operator: It was established that the reason for problem 3) is that C is a scalar densityof weight two while it was shown that only density weight one scalars have a chance to be quantizedrigorously and background independently. Therefore the currently accepted point of view is that βshould be real and that one uses the original unrescaled C rather than C.

III.1.1 A Possible New Mechanism for Avoiding UV Singularities inBackground Independent Quantum Field Theories

Before we go into more details concerning [48], let us give a heuristic explanation just why it happensthat QGR may cure UV problems of QFT, making the connection with the issue of the densityweight just mentioned. Consider classical Einstein – Maxwell theory on M = R × σ in its canonicalformulation, then the Hamiltonian constraint gains an extra matter piece given for unit lapse N = 1by

HEM =1

2e2

σ

d3x

Density weight -1︷ ︸︸ ︷qab√det(q)

[EaEb +BaBb]︸ ︷︷ ︸Density weight +2

(III.1.1.1)

Exercise III.1.1.Starting from the Lagrangean

L = − 1

4e2

√| det(g)|FµνFρσgµρgνσ (III.1.1.2)

46

where F = 2dA is the spacetime curvature of the Maxwell connection A with unit cm−1 and e is theelectric charge in units such that α = ~e2 is the dimensionfree Feinstrukturkonstante, perform theLegendre transform. With the electric field Ea being the momentum conjugate to the spatial piece Aaof A verify that the “Hamiltonian” is given by −A0G+NaV ′

a+NC′ where G = ∂aE

a is the Gauss law,V ′a = FabE

b and C ′ is the integrand of (III.1.1.1) with Ba = ǫabcFbc/2 the magnetic field. Check thatG′ generates U(1) gauge transformations while V ′

a generates spatial diffeomorphisms where Aa, Ea

transform as a one – form and a vector density of weight one respectively. Conirm that also Ba is avector density of weight one.

As the exercise reveals, the geometry factor in (III.1.1.1) is a symmetric covariant tensor of ranktwo of density weight −1 due to the factor

√det(q) in the denominator while the matter part is a

symmetric contravariant tensor of rank two of density weight +2. That the resulting scalar has netdensity weight is +1 is no coincidence but a direct consequence of the diffeomorphism invarianceor background independence of any matter theory coupled to gravity: only the integral over σ of ascalar density of weight +1 is spatially diffeomorphism invariant.

We can now quantize (III.1.1.1) in two ways:1)In the first version we notice that if g = η is the Minkowski metric, that is, qab = δab then (III.1.1.1)reduces to the ordinary Maxwell Hamiltonian on Minkowski space. Thus we apply the formalismof QFT on a background spacetime, in this case Minkowski space, because we have fixed qab to thenon-dynamical C−number field δab which is not quantized at all.2)In the second version we keep qab dynamical and quantize it as well. Thus we apply QGR, abackground independent quantization. Now qab becomes a field operator qab and the statementthat the metric is flat can at most have a semiclassical meaning, that is, the expectation value of qabin a gravitational state is close to δab.Let us now sketch how these two different quantizations are performed and exactly pin-point how ithappens that the first quantization is divergent while the second is finite.

1) QFT on a background spacetimeAs we have said, the metric qab = δab is now no longer a dynamical entitity but just becomes acomplex number. What we get is the usual Maxwell Hamiltonian operator

HM =1

2e2

Σ

d3x δab[EaEb + BaBb] (III.1.1.3)

Notice the crucial difference with (III.1.1.1): The net density weight of the operator valueddistribution in the integral is now +2 rather than +1 ! By switching off the metric as adynamical field we have done a severe crime to the operator, because the net density weight +2will be remembered by the operator in any faithful representation of the canonical commutationrelations and leads to the following problem: The only coordinate density of weight one thatone can construct is a δ−distribution (and derivatives thereof), thus for instance the operatorEa(x) is usually represented as a functional derivative which one can rewrite formally as

αδ/δAa(x) = α∑

y∈Σ

δ(x, y)∂/∂Aa(y) (III.1.1.4)

The right hand side of (III.1.1.4) is a sum over terms each of which consists of a well-definedoperator Ya(y) = ∂/∂Aa(y) and a distributional prefactor δ(x, y). It is for this reason that

47

expressions of the form Ea(x)Eb(x) cannot be well-defined since we get products of distribu-tions supported at the same point x and which result in divergent expressions of the form∑

y,z δ(x, y)δ(x, z)Ya(y)Yb(z) =∑

y δ(x, y)2Ya(y)Yb(y). The density weight two is correctly en-

coded in the term δ(x, y)2 = δ(0, 0)δ(x, y) which, however, is meaningless.

These heuristic arguments can of course be made precise: (III.1.1.3) is quantized on the Fockspace HFock and one obtains

HM = : HM : + ~

Σ

[√

−∆xδ(x, y)]x=y︸ ︷︷ ︸UV Singularity

(III.1.1.5)

Here the colons stand for normal ordering. The UV (or short distance) singularity is explicitlyidentified as the coincidence limit x = y of the integrand in the normal ordering correction.Therefore HM is ill-defined on HFock. Notice that even if the integrand would be finite, theintegral suffers from an IR (or large volume) singularity if σ is not compact which comes fromthe fact that we are dealing with an infinite number of degrees of freedom. This singularity is,in contrast to the UV singularity, physical since it captures the vacuum energy of the universewhich is of course infinite if the volume is.

2) QFT coupled to QGRThis time we keep the metric as a dynamical variable and quantize it. Thus instead of (III.1.1.3)we obtain something of the form

HEM =1

2e2

Σ

d3xqab√det(q)

[EaEb + BaBb] (III.1.1.6)

This time the net density weight is still +1. Now while the expression (III.1.1.4) is still validand implies that there will be a product of δ−distributions in the numerator coming from thematter operator valued distributions, there is also a δ−distribution in the denominator due tothe factor

√det(q) which comes about as follows: As we already mentioned in section II.2.2

the volume functional in (II.2.2.4) admits a well-defined quantization of the form

Vol(R)Ts = ℓ3P∑

v∈V (γ(s))∩RVvTs (III.1.1.7)

where Vv is a well-defined, dimensionfree operator (not an operator valued distribution !) builtfrom the vector fields vjS. Since Vol(R) is the integral over R of

√det(q) we conclude that√

det(q) admits a quantization as an operator valued distribution, namely

√det(q)(x)Ts = ℓ3P

v∈V (γ(s))

δ(x, v)VvTs (III.1.1.8)

Now certainly (III.1.1.6) cannot be quantized on the Hilbert space Hkin⊗HFock because HFock

depends on a background metric (for instance through the Laplacian ∆) which is not availableto us. However, we may construct a background independent Hilbert space H′

kin for Maxwelltheory which is completely identical to our Hkin, just that SU(2) is replaced by U(1) [48].In H′

kin the role of spin network states is played by charge network (CNW) states, that is,edges e are labelled by integers ne (irreducibles of U(1). Let us denote CNW’s by c = (γ, ~n =

48

nee∈E(γ)) and CNWF’s by T ′c. Then a basis for the Einstein-Maxwell theory kinematical

Hilbert space Hkin ×H′kin is given by the states Ts ⊗ T ′

c.

Now something very beautiful happens, which is not put in by hand but rather is a derivedresult: A priori the states Ts, T

′c may live on different graphs, however, unless the graphs are

identical, the operator automatically (III.1.1.6) annihilates Ts ⊗ T ′c [49]. This is the mathe-

matical manifestation of the following deep physical statement: Matter can only exist wheregeometry is excited. Indeed, if we have a gravitational state which has no excitations in acoordinate region R then the volume of that region as measured by the volume operator isidentically zero. However, if a coordinate region has zero volume, then it is physically sim-ply not there, it is empty space. Summarizing, the operator (III.1.1.6) is non-trivial only ifγ(s) = γ(c).

With this being understood, let us then sketch the action of (III.1.1.6) on our basis. One findsheuristically

HEMTs ⊗ T ′c = mP

v∈V (γ)

e,e′∈E(γ),e∩e′=v×

×∫

Σ

d3x[qe,e′

Vv

1

δ(x, v)︸ ︷︷ ︸↑

Ts] ⊗ [δ(x, v)︸ ︷︷ ︸↑

︸ ︷︷ ︸Cancellation

δ(y, v)Y eY e′T ′c]x=y] (III.1.1.9)

where mP =√

~/κ is the Planck mass. Here qe,e′ and Y e are well-defined, dimensionfreeoperators (not distribution valued !) on Hkin and H′

kin respectively built from the right invariantvector fields Rj

e, Re that enter the definition of the flux operators as in (II.1.1.12) and its analogfor U(1). The product of δ−distributions in the numerator of (III.1.1.9) has its origin again inthe fact that the matter operator has density weight +2 certainly also in this representationand therefore has to be there, so nothing is swept under the rug! The δ−distribution inthe denominator comes from (III.1.1.8) and correctly accounts for the fact that the geometryoperator has density weight −1. Again we have a coincidence limit x = y which comes from apoint splitting regularization and which in the background dependent quantization gave rise tothe UV singularity. Now we see what happens: One of the δ−distributions in the numeratorgets precisely cancelled by the one in the denominator leaving us with only one δ−distributioncorrectly accounting for the fact that the net density weight is +1. The integrand is thenwell-defined and the integral can be performed resulting in the finite expression

HEMTs ⊗ T ′c =

v∈V (γ)

e,e′∈E(γ),e∩e′=v×

× [qe,e′

VvTs] ⊗ [Y eY e′T ′

c]x=y (III.1.1.10)

Notice that finite here means non-perturbatively finite, that is, not only finite order by orderin perturbation theory (notice that in coupling gravity we have a highly interacting theory infront of us). Thus, comparing our non-perturbative result to perturbation theory the resultobtained is comparable to showing that the perturbation series converges ! Notice also thatfor non-compact σ the expression (III.1.1.10) possibly has the physically correct IR divergencecoming from a sum over an infinite number of vertices.

49

Exercise III.1.2.Recall the Fock space quantization of the Maxwell field and verify (III.1.1.5).

This ends our heuristic discussion about the origin of UV finiteness in QGR. The crucial point isobviously the density weight of the operator in question which should be precisely +1 in order toarrive at a well-defined, background independent result: Higher density weight obviously leads tomore and more divergent expressions, lower density weight ends in zero operators.

III.1.2 Sketch of a Possible Quantization of the Hamiltonian Constraint

We now understand intuitively why the rescaled Hamiltonian constraint C had no chance to bewell-defined in the quantum theory: It is similar to (III.1.1.5) due to its density weight +2. Thesame factor 1/

√det(q) that was responsible for making (III.1.1.6) finite also makes the original,

non-polynomial, unrescaled Hamiltonian constraint C = C/√

det(q) finite. We will now proceed tosome details how this is done, avoiding intermediate divergent expressions such as in (III.1.1.9).

The essential steps can already be explained for the first term in (I.2.2.5) so let us drop the“More” term and consider only the integrated first term

CE(N) =1

κ

Σ

d3xNF jabǫjklE

ajE

bl√

| det(E)|(III.1.2.1)

Let us introduce a mapR : Σ → O(Σ); x 7→ Rx (III.1.2.2)

where O(Σ) denotes the set of open, compactly supported, connected and simply connected subsetsof Σ and Rx ∈ O(Σ) is constrained by the requirement that x ∈ Rx. We define the volumes of theRx by

V (x) := Vol(Rx) =

Rx

d3y√| det(E)|(y) (III.1.2.3)

Then, up to a numerical prefactor we may write (III.1.2.1) in the language of differential forms andin terms of a Poisson bracket as

CE(N) =1

κ2

Σ

NTr(F ∧ A, V ) (III.1.2.4)

Exercise III.1.3.Verify that (III.1.2.3) is really the volume of Rx and (III.1.2.4).

The reasoning behind (III.1.2.4) was to move the factor 1/√

det(q)(x) from the denominator intothe numerator by using a Poisson bracket. This will avoid the δ−distribution in the denominator asin (III.1.1.9) and has the additional advantage that

√det(q) now appears smeared over Rx so that

one obtains an operator, not a distribution. Thus, the idea is now to replace the function V (x) by

the well-defined operator Vol(Rx) and the Poisson bracket by a commutator divided by i~. The onlything that prevents us from doing this is that the operators Aja, F

jab do not exist on Hkin. However,

they can be regularized in terms of holonomies as follows:Given tangent vectors u, v ∈ Tx(Σ) we define one parameter homotopies of paths and loops of triangletopology

ǫ 7→ puǫ,x, αuvǫ,x (III.1.2.5)

50

respectively with b(puǫ,x) = b(αuvǫ,x) = x and (puǫ,x)x = (αuvǫ,x)x+ = ǫu, (αuvǫ,x)x− = −ǫv (left and rightderivatives at x). Then for smooth connections A ∈ A the Ambrose – Singer theorem tells us that

limǫ→0

[A(puǫ,x − 1]/ǫ = uaAa(x), limǫ→0

[A(αuvǫ,x − 1]/ǫ2 = uavbFab(x)/2 (III.1.2.6)

Exercise III.1.4. Verify (III.1.2.6) by elementary means, using directly the differential equation(II.1.1.1).Hint:For sufficiently small ǫ we have up to ǫ2 corrections puǫ,x(t) = x+ ǫtu and

αuvǫ,x(t) = x+ ǫ

tu t ∈ [0, 1]u/3 + (t− 1)(v − u) t ∈ [1, 2]

(3 − t)v t ∈ [2, 3]

Thus, given a triangulation τǫ of σ, that is, a decomposition of σ into tetrahedra ∆ with basepoints v(∆), edges pI(∆), I = 1, 2, 3 of ∆ of the type puǫ,v starting at v and triangular loops αIJ(∆) =pI(∆) aIJ(∆)pJ(∆)−1 of the type αuvǫ,v where the arcs aIJ(∆) comprise the remaining three edges of∆, it is easy to show, using (III.1.2.6), that up to a numerical factor

CτǫE (N) =

1

κ2

∆∈τǫN(v(∆))

IJK

ǫIJKTr(A(αIJ(∆)A(pK(∆))A(pK(∆))−1, V (v) (III.1.2.7)

tends to CE(N) as ǫ→ 0 (in this limit the triangulation gets finer and finer).

Exercise III.1.5.Verify this statement.

The expression (III.1.2.7) can now be readily quantized on Hkin because holonomies and volumefunctionals are well-defined operators. However, we must remove the regulator ǫ in order to arrive ata quantization of (III.1.2.4). Now the regulator can be removed in many inequivalent ways becausethere is no unique way to refine a triangulation. Moreover, we must specify in which operatortopology Cτǫ(N) converges. The discussion of these issues is very complicated and the interestedreader is referred to [48] for the detailed arguments that lead to the following solution:

i) TriangulationFirst of all we define the operator explicitly on the SNW basis Ts. In order for the refinementlimit to be non-trivial, it turns out that the triangulation must be refined in such a way thatγ(s) ⊂ τǫ for sufficiently small ǫ. This happens essentially due to the volume operator whichhas non-trivial action only at vertices of graphs. Thus the refinement must be chosen dependingon s. This is justified because classically all refinements lead to the same limit. One mightworry that this does not lead to a linear operator, however, this is not the case because it isdefined on a basis.

ii) Operator TopologyThe limit ǫ→ 0 exists in the following sense:Let D∗

Diff ⊂ D∗kin be the space of solutions of the diffeomorphism constraint. We say that a fam-

ily of operators Oǫ converges to an operator O on Hkin in the uniform – weak – Diff∗−topologyprovided that for each δ > 0 and for each l ∈ D∗

Diff , f ∈ Dkin there exists ǫ(δ) > 0 independentof l, f such that

|l([Oǫ − O]f)| < δ ∀ǫ < ǫ(δ) (III.1.2.8)

51

v,e’)

a(γ ,v,e,e’)

α(γ,v,e,e’)

v

e

e’

e’’

γ,p(

Figure 11: Meaning of the loop, path and arc assignment of the Hamiltonian constraint. Notice howa tetrahedron emerges from those objects, making the link with the triangulation. The broken linesindicate possible other edges or continuations thereof.

This topology is of course motivated by physical considerations: Since the operator is un-bounded, the uniform (i.e. operator norm) topology is too strong. The strong or weak topolo-gies (pointwise convergence in Hilbert space norm or as matrix elements) give a trivial (zero)limit (excercise !). Thus one is naturally led to ∗ topologies. The maximal dual space onwhich to build a topology would be D∗

kin but one can check that the limit does not exist evenpointwise in D∗

kin. Thus one is looking for suitable subspaces thereof. The natural, physicallymotivated choice is, of course, the space D∗

Diff which is singled out by the spatial diffeomor-phism constraint. The reason for why have required uniform convergence in (III.1.2.8) is thatthis excludes the existence of the limit for larger spaces D∗

Diff ⊂ D∗⋆ ⊂ D∗

kin.

The end result is

C†E(N) = mP

v∈V (γ(s))

N(v)∑

e, e′, e′′ ∈ E(γ(s))e ∩ e′ ∩ e′′ = v

×

× Tr([A(αγ(s),v,e,e′) − (A(αγ(s),v,e,e′))−1]A(pγ(s),v,e′′)[A(pγ(s),v,e′′)

−1, Vv])

+cyclic permutation ine, e′, e′′ Ts (III.1.2.9)

The meaning of the loops αγ,v,e,e′ and paths pγ,v,e′′ that appears in this sum over vertices and triples ofedges incident at them is maybe best explained in the following figure 11. Their precise specificationmakes use of the axiom of choice and is diffeomorphism covariant, that is, for ϕ ∈ Diffω(Σ), e.g. theloops αγ,v,e,e′ and αϕ(γ),ϕ(v),ϕ(e),ϕ(e′) are analytically diffeomorphic. Moreover, the arcs aγ,v,e,e′ definedby

αγ,v,e,e′ = pγ,v,e aγ,v,e,e′ p−1γ,v,e′ (III.1.2.10)

are such that also γ ∪ aγ,v,e,e′ and ϕ(γ) ∪ aϕ(γ),ϕ(v),ϕ(e),ϕ(e′) are anlytically diffeomorphic. The adjoint

in (III.1.2.9) is due to the fact that CE(N) is classically real-valued, so we are quantizing CE(N) aswell. The operator (III.1.2.9) is not symmetric, however, its adjoint is densely defined on Dkin andit is therefore closable. Usually one requires real valued functions to become self-adjoint operators

52

because then by the spectral theorem the spectrum (possible measurement values) is a subset of thereal line. However, this argument is void when we are only interested in the kernel of the operator(“zero eigenvalue”).

Exercise III.1.6.Verify that C†

E(N) is not symmetric but it is, together with CE(N), densely defined on Dkin. Showthat if real valued constraints CI form a Poisson algebra CI , CJ = fIJ

KCK with non-trivial,real valued structure functions such that fIJ K , CKCL=0 6= 0, then CI , fIJ

K must not be bothsymmetric in order for the quantum algebra to be free of anomalies. Conclude that the failure of(III.1.2.9) to be symmetric is likely to be required for reasons of consistency.

The fact the loop αγ,v,e,e′ is not shrunk to v as one would expect is of course due to our definitionof convergence, in fact, an arbitrary loop assignment (γ, v, e, e′) 7→ αγ,v,e,e′ that has the same dif-feomorphism invariant characteristics is allowed, again because in a diffeomorphism invariant theorythere is no notion of “closeness” of αγ,v,e,e′ to v. Notice that the operator CE(N) is defined on Hkin

using the axiom of choice and not on diffeomorphism invariant states as it is sometimes misleadinglystated in the literature [50]. In fact, it cannot be because the dual operator C ′

E(N) defined by

[C ′E(N)l](f) := l(C†

E(N)f) (III.1.2.11)

for all f ∈ Dkin, l ∈ D∗kin does not preserve D∗

Diff as is expected from the the classical Poisson algebraV, C ∝ C 6= V . If one wants to take this dual point of view then one is forced to introduce a largerspace D∗

⋆ which is preserved but which does not solve the diffeomorphism constraint and is thereforeunphysical. This has unnecessarily given rise to a large amount of confusion in the literature andshould be abandomed.

As we have said, the loop assignment is to a very large extent arbitrary at the level of Hkin

and represents a serious quantization ambiguity, it cannot even be specified precisely because weare using the axiom of choice. However, at the level of Hphys this ambiguity evaporates to a largeextent because all choices that are related by a diffeomorphism result in the same solution space toall constraints defined by elements l ∈ D∗

Diff which satisfy in addition

[C ′E(N)l](f) = l(C†

E(N)f) = 0 ∀ N ∈ C∞0 (Σ), f ∈ Dkin (III.1.2.12)

where C∞0 (Σ) denotes the smooth functions of compact support. Thus the solution space D∗

phys

will depend only on the spatially diffeomorphism invariant characteristics of the loop assignmentwhich can be specified precisely [48], it essentially characterizes the amount by which the arcs knotthe original edges of the graph. Besides this remaining ambiguity there are also factor orderingambiguities but no singularities some of which are discussed in [51].

Let us list without proof some of the properties of this operator:

i) Matter CouplingSimilar Techniques can be applied to the case of (possibly supersymmetric) matter coupled toGR [48].

ii) Anomaly – FreenessThe constraint algebra of the Hamiltonian constraint with the spatial diffeomorphism con-straint and among each other is mathematically consistent. From the classical constraint al-gebra V, C ∝ C we expect that V (ϕ)C†

E(N)V (ϕ)−1 = CE(ϕ∗N) for all diffeomorphisms ϕ.However, this is just the statement of the loop assignment being diffeomorphism covariant

53

which can be achieved by making use of the axiom of choice. Next, from C,C ∝ V we ex-pect that the dual of [C†

E(N), C†E(N ′)] = [CE(N ′), CE(N)]† annihilates the elements of D∗

Diff .

This can be explicitly verified [48]. We stress that [C†E(N), C†

E(N ′)] is not zero, the algebra ofHamiltonian constraints is not Abelean as it is sometimes misleadingly stated in the literature.The commutator is in fact explicitly proportional to a diffeomorphism.

iii) Physical StatesThere is a rich space of rigorous solutions to (III.1.2.12) and a precise algorithm for theirconstruction has been developed [48].

iv) Intuitive PictureThe Hamiltonian constraint acts by annihilating and creating spin degrees of freedom andtherefore the dynamical theory obtained could be called “Quantum Spin Dynamics (QSD)” inanalogy to “Quantum Chromodynamics (QCD)” in which the Hamiltonian acts by creatingand annihilating colour degrees of freedom. In fact we could draw a crude analogy to Fockspace terminology as follows: The (perturbative) excitations of QCD carry a continuous label,the mode number k ∈ R3 and a discrete label, the occupation number n ∈ N (and others). InQSD the continuous labels are the edges e and the discrete ones are spins j (and others). Sowe have something like a non – linear Fock representation in front of us.

Next, when solving the Hamiltonian constraint, that is, when integrating the Quantum EinsteinEquations, one realizes that one is not dealing with a (functional) partial differential equationbut rather with a (functional) partial difference equation. Therefore, when understandingcoordinate time as measured how for instance volumes change, we conclude that also timeevolution is necessarily discrete. Such discrete time evolution steps driven by the Hamiltonianconstraint assemble themselves into what nowadays is known as a spin foam. A spin foam is afour dimensional complex of two dimensional surfaces where each surface is to be thought of asthe world sheet of an edge of a SNW and it carries the spin that the edge was carrying beforeit was evolved7.

Another way of saying this is that a spin foam is a complex of two-surfaces labelled by spinsand when cutting a spin foam with a spatial three-surface Σ one obtains a SNW. If one uses twosuch surfaces Σt,Σt+TP

where TP = ℓP/c is the Planck time then one rediscovers the discretetime evolution of the Hamiltonian constraint. These words are summarized in figure 12.

While these facts constitute a promising hint that the Hilbert space Hkin could in fact support thequantum dynamics of GR, there are well-taken concerns about the physical correctness of the oper-ator C†

E(N):

The problem is that one would like to see more than that the commutator of two dual Hamilto-nian constraints annihilates diffeomorphism invariant states, one would like to see something of thekind

[C†E(N), C†

E(N ′)] = iℓ2P

[

Σ

d3x[NN ′,a −N,aN ′]qabVb] (III.1.2.13)

The reason for this is that then one would be sure that C†E(N) generates the correct quantum

evolution. While this requirement is not necessary, it is certainly sufficient and would be reassuring8.

7Thus, a spin foam model can be thought of as a background independent string theory !8Example: Suppose that Ca are the angular momentum components for a particle in in R3 with classical constraint

54

Figure 12: Emergence of a spin foam from a SNW by the action of the Hamiltonian constraint

There are two obstacles that prevent us from rewriting the left hand side of (III.1.2.13) in terms ofthe right hand side.1)The one parameter groups s 7→ V (ϕus ) of unitarities where ϕus are the one parameter groups ofdiffeomorphisms defined by the integral curves of a vector field u are not weakly continuous, thereforea self-adjoint generator V (u) that we would like to see on the right hand side of (III.1.2.13) simplydoes not exist.

Exercise III.1.7.Recall Stone’s theorem about the existence of the self-adjoint generators of weakly continuous one-parameter unitary groups and verify that V (ϕus ) is not weakly continuous on Hkin.

2)One can quantize the right hand side of (III.1.2.13) by independent means and it does annilate D∗

Diff

[48], however, that operator does not resemble the left hand side in any obvious way. The reasonfor this is that even classically it takes a DinA4 page of calculation in order to rewrite the Poissonbracket CE(N), CE(N ′) as in (III.1.2.13) with Va given by (I.2.2.5). The manipulations that mustbe performed in order to massage the Poisson bracket into the desired form involve a) integrations bypart, b) writing Fab in terms of Aa, c) derivatives of

√det(q), d) multiplying fractions by functions

in both numerator and denominator, e) symmetry arguments in order to see that certain tems canceletc. (exercise !). These steps are obviously difficult to perform with operators.

In summary, there is no mathematical inconsistency, however, there are doubts about the physi-cal correctness of the Hamiltonian constraint operator presently proposed although no proof existsso far that it is necessarily wrong. In order to make progress on this issue, it seems that we needto develop first a semiclassical calculus for the theory, more precisely, we need coherent states sothat expectation values of operators and their commutators can be replaced, up to ~ corrections, bytheir classical values and Poisson brackets respectively for which then the manipulations listed in 2)

algebra Ca, Cb = ǫabcCc. Introduce polar coordinates and define the non-self adjoint operators C1 = i~∂/∂θ, C2 =i~∂/∂φ, C3 = 0. Then the quantum constraint algebra is Abelean and does not at all resemble the classical one,however, the physical states are certainly the correct ones, functions that depend only on the radial coordinate.

55

above can be carried out. If that is possible, and the outcome of these calculations is the expectedone, possibly after changing the operator by making use of the available quantization ambiguities,then one would be able to claim that one has indeed constructed a quantum theory of GR with thecorrect classical limit. Only then can one proceed to solve the theory, that is, to construct solutions,the physical inner product and the Dirac observables. The development of a semiclassical calculusis therefore one of the “hot” research topics at the moment.

Another way to get confidence in the quantization method applied to the Hamiltonian constraintis to study model systems for which the answer is known. This has been done for 2+1 gravity [48]and for quantum cosmology to which we turn in the next section.

III.2 Loop Quantum Cosmology

III.2.1 A New Approach To Quantum Cosmology

The traditional approach to quantum cosmology consists in a so-called mini – superspace quantiza-tion, that is, one imposes certain spacetime Killing symmetries on the metric, plugs the symmetricmetric into the Einstein Hilbert action and obtains an effective action which depends only on a finitenumber of degrees of freedom. Then one canonically quantizes this action. Thus one symmetrizesbefore quantization. These models are of constant interest and have natural connections to inflation.See e.g. [52] for recent reviews.

What is not perfect about these models is that 1) not only do they switch off all but an infinitenumber of degrees of freedom, but 2) also the quantization method applied to the reduced modelusually is quite independent from that applied to the full theory. A fundamental approach to quantumcosmology will be within the full theory and presumably involves the construction of semiclassicalphysical states whose probability amplitude is concentrated on, say a Friedmann – Robertson –Walker (FRW) universe. This would cure both drawbacks 1) and 2). At the moment we cannotreally carry out such a programme since the construction of the full theory is not yet complete.However, one can take a more modest, hybrid approach, where while dealing only with a finitenumber of degrees of freedom one takes over all the quantization machinery from the full theory !Roughly speaking, one works on the space Hkin of the full theory but considers only states thereinwhich satisfy the Killing symmetry. Hence one symmetrizes after quantization which amounts toconsidering only a finite subset of holonomies and fluxes. This has the advantage of leading to asolvable model while preserving pivotal structures of the full theory, e.g. the volume operator appliedto symmetric states will still have a discrete spectrum as in the full theory while in the traditionalapproaches it is continuous. Such a programme has been carried out in great detail by Bojowaldin a remarkable series of papers [53] and his findings are indeed spectacular, should they extend tothe full theory: It turns out that the details of the quantum theory are drastically different from thetraditional minisuperspace approach. In what follows we will briefly describe some of these results,skipping many of the technical details.

III.2.2 Spectacular Results

Consider the FRW line element (in suitable coordinates)

ds2 = −dt2 +R(t)2[dr2

1 − kr2+ r2dΩ2

2] =: −dt2 + R(t)2q0abdx

adxb (III.2.2.1)

56

Figure 13: Spectrum of the inverse scale factor

The universe is closed/flat/open for k = 1/0/ − 1. The only dynamical degree of freedom left isthe so-called scale factor R(t) which describes the size of the universe and its conjugate momentum.The classical big bang singularity corresponds to the fact that the Einstein equations predict thatlimt→0R(t) = 0 at which the metric (III.2.2.1) becomes singular and the inverse scale factor 1/R(t)blows up (the curvature will be ∝ 1/R(t)2 so this singularity is a true curvature singularity).

We are interested in whether the curvature singularity 1/R → ∞ exists also in the quantumtheory. To study this we notice that for (III.1.1.3) det(q) = R6 det(q0). Hence, up to a numericalfactor this question is equivalent to the question whether the operator corresponding to 1/ 6

√det(q),

when applied to symmetric states, is singular or not. However, we saw in the previous section thatone can trade a negative power of det(q) by a Poisson bracket with the volume operator. In [53]precisely this, for the Hamiltonian constraint, essential quantization technique is applied which iswhy this model tests some aspects of the quantization of the Hamiltonian constraint. Now it turns

out that this operator, applied to symmetric states, leads to an operator 1R

which is diagonalized by(symmetric) SNWF’s and the spectrum is bounded ! In figure 13 we plot the qualitative behaviourof the eigenvalues ℓPλj as a function of j where j is the spin label of a gauge invariant SNWF witha graph consisting of one loop only (that only such states are left follows from a systematic analysiswhich defines what a symmetric SNWF is). One can also quantize the operator R and one sees thatits eigenvalues are essentially given by jℓP up to a numerical factor. Thus the classical singularitycorresponds to j = 0 and one expects the points λjℓp at the values 1/j on the curve ℓP/R. Evidentlythe spectrum is discrete (pure point) and bounded, at the classical singularity it is finite. In otherwords, the quantum universe never decreases to zero size. For larger j, in fact already for R of theorder of ten Planck lengths and above, the spectrum follows the classical curve rather closely hintingat a well-behaved classical limit (correspondence principle).

Even more is true: One can in fact quantize the Hamiltonian constraint by the methods of theprevious section and solve it exactly. One obtains an eighth order difference equation (in j). Thesolution therefore depends non-trivially on the initial condition. What is surprising, however, is thefact that only one set of initial conditions leads to the correct classical limit, thus in loop quantumcosmology initial conditions are derived rather than guessed. One can even propagate the quantumEinstein equations through the classical singularity and arrives at the picture of a bouncing universe.

Finally one may wonder whether these results are qualitatively affected by the operator ordering

57

ambiguities of the Hamiltonian constraint. First of all one finds that these results hold only if oneorders the loop in (III.1.2.9) to the left of the volume operator as written there. However, one isnot forced to work with the holonomy around that loop in the fundamental representation of SU(2),there is some flexibility [51] and one can choose a different one, say j0. It turns out that the valuej0 influences the onset of classical behaviour, that is, the higher j0 the higher the value j(j0) fromwhich on the spectrum in figure 13 lies on the curve 1/j. Now this is important when one copuples,

say scalar matter because the operator 1R

enters the matter part of the Hamiltonian constraint andmodifies the resulting effective equation for R(t) in the very early phase of the universe and leads toa quantum gravity driven inflationary period whose duration gets larger with larger j0 !

Thus, loop quantum cosmology not only confirms aspects of the quantization of the Hamiltonianconstraint but also predicts astonishing deviations from standard quantum cosmology which oneshould rederive in the full theory.

III.3 Path Integral Formulation: Spin Foam Models

III.3.1 Spin Foams from the Canonical Theory

Spin Foam models are the fusion of ideas from topological quantum field theories and loop quantumgravity, see e.g. [54] for a review, especially the latest, most updated one by Perez. The idea thatconnects these theories is actually quite simple to explain at an heuristic level:

If we forget about 1) all functional analytic details, 2) the fact that the operator valued distri-butions corresponding to the Hamiltonian constraint C(x) do not mutually commute for differentx ∈ σ and 3) that the Hamiltonian constraint operators C(N) are certainly not self-adjoint, at leastas presently formulated, then we can formally write down the complete space of solutions to theHamiltonian constraint as a so-called “rigging map” (see e.g. [1])

η : Dkin → D∗phys; f 7→ δ[C] f := [

x∈Σ

δ(C(x)) f ] (III.3.1.1)

(where η = c.c · η is the complex conjugate of the actual anti-linear rigging map). Here theδ−distribution of an operator is defined via the spectral theorem (assuming the operator to beself-adjoint). Notice that we do not need to order the points x ∈ σ as we assumed the C(x) to bemutually commuting for the moment and only under this assumption it is true that, at least for-mally η[f ](C(N)f ′) = 0 (exercise)9. Now we use the formula δ(x) =

∫Rdk2πeikx to write the functional

δ−distribution δ[C] as a path integral

δ[C] =

N ′

[DN ]eiC(N) (III.3.1.2)

where we have neglected an infinite constant as usual in this formal business. Here N ′ is the spaceof lapse functions at a fixed time. Let us introduce also the space of lapses with arbitray timedependence Nt1,t2 in t ∈ [t1, t2]. Then, up to an infinite constant one can verify that

δ[C] =

N t2t1

[DN ]ei∫ t2t1dt

∫Σd3xN(x,t)C(x) (III.3.1.3)

9At an even more formal level η[f ] is also a solution in the non-commuting case if, as is the case with the currentlyproposed C, the algebra with the spatial diffeomorphism constraint closes

58

The rigging map machinery then tells us that the scalar product on the image of the rigging map issimply given by

< η(f), η(f ′) >phys:=< f, η(f ′) >kin=

N t2t1

[DN ] < f, ei∫ t2

t1dtC(Nt)f ′ >kin (III.3.1.4)

This formula looks like a propagator formula, that is, like a transition amplitude between an initialstate f ′ on Σt1 and a final state f on Σt2 after a multi-fingered time evolution generated by C(Nt).In fact, if we use the Taylor expansion of the exponential function and somehow regularize thepath integral then the expansion coefficients < Ts, C(Nt)

nT ′s′ >kin can be interpreted as probabilityamplitude of the evolution of the SNW state T ′

s′ to reach the SNW state Ts after n time steps (recallfigure 12).

Now by the usual formal manipulations that allow us to express a unitary operator ei(t2−t1)H asa path integral over the classical pase space M (the rigorous version of which is the Feynman – Kacformula, e.g. [56]) one can rewrite (III.3.1.4) as

< η(f), η(f ′) >phys=

∫[DN D ~N DΛ DA DE] < f, eiSf ′ >kin (III.3.1.5)

where S is the Einstein-Hilbert action written in canonical form in terms of the variables A,E, thatis

S =

R

dt

Σ

d3xAjaEaj − [−ΛjGj +NaVa +NC] (III.3.1.6)

and we have simultaneously included also projections on the space of solutions to the Gauss andvector constraint. Now the action (III.3.1.6) is the 3 + 1 split of the following covariant action

S =

M

ΩIJ ∧ [ǫIJKL − β−1ηIKηJK]eK ∧ eL (III.3.1.7)

discovered in [57] where β is the Immirzi parameter. Here ΩIJ is the (antisymmetric) curvaturetwo-form of an (antisymmetric) SL(2,C) connection one-form ωIJ with Lorentz indices I, J,K, .. =0, 1, 2, 3, η is the Minkowski metric and eI is the co-tetrad one-form. The first term in (III.3.1.7)is called the Palatini action while the second term is topological (a total differential modulo theequations of motion). The relationn between the four-dimensional fields ωIJµ , e

Iµ (40 components)

and the three-dimensional fields Aja, Eaj ,Λ

j, N,Na (25 components) can only be established if certainso-called second class constraints [20] are solved.

III.3.2 Spin Foams and BF – Theory

Thus, it is formally possible to write the inner product between physical states as a covariant pathintegral for the classical canonical action and using only the kinematical inner product, thus providinga bridge between the covariant and canonical formalism. However, this bridge is far from beingrigorously established as we had to perform many formal, unjustified manipulations. Now ratherthan justifying the steps that lead from C to (III.3.1.5) one can turn the logic upside down and startfrom a manifestly covariant formulation and derive the canonical formultion. This is the attitudethat one takes among people working actively on spin foam models. Thus, let us forget about thetopological term in (III.3.1.7) and consider only the Palatini term. Then the Palatini action hasprecisely the form of a BF – action

SBF =

M

ΩIJ ∧BIJ (III.3.2.1)

59

just that the (antisymmetric) two-form field BIJ is not arbitrary (it would have 36 independentcomponents), it has to come from a tetrad with only 16 independent components, that is, it has tobe of the form ǫIJKLeK ∧ eL.

Exercise III.3.1.Show that the condition that B comes from a tetrad is almost10 equivalent to the simplicity constraint

ǫIJKLBIJµνB

KLρσ = cǫµνρσ (III.3.2.2)

for some spacetime scalar density c of weight one.

The reasoning is now as follows: BF – theory without the constraint (III.3.2.2) is a topologicalfield theory, that is, it has no local degrees of freedom. Therefore quantum BF – theory is not reallya QFT but actually a quantum mechanical system and can therefore be handled much more easilythan gravity. Let us now write an action equivalent to the Palatini action given by

S ′P [ω,B,Φ] = SBF [ω,B] + SI [B,Φ]

SI [B,Φ] :=

M

ΦµνρσǫIJKLBIJαβB

KLγδ [δαµδ

βν δ

γρδ

δσ −

1

4!ǫαβγδǫµνρσ] (III.3.2.3)

where the Lagrange multiplier Φµνρσ [58] is a four dimensional tensor density of weight one, symmetricin the index pairs (µν) and (ρσ) and antisymmetric in each index pair. Thus, Φ has (6 · 7)/2 = 21independent components of which the totally skew component is projected out in (III.3.2.3), leavingus with 36 − 16 = 20 independent components. Hence the Euler Lagrange equations for Φ preciselydelete the amount of unwanted degrees of freedom in B and impose the simplicity constraint. Hence,classically S ′

P [Ω, B,Φ] and SP [Ω, e] are equivalent. Thus, if we write a path integral for S ′P and treat

the Lagrange multiplier term SI in (III.3.2.3) as an interaction Lagrangean (a perturbation) to BF– theory, then we can make use of the powerful techniques that have been developed for the pathintegral quantization for BF – theory and its perturbation theory.

Exercise III.3.2.i) Write the Euler Lagrange equations for BF – theory and conclude that the solutions consist of flatconnections ω and gauge invariant B− fields. Conclude that ω can be gauged to zero by SL(2,C)transformations locally and that then B is closed, that is, locally exact by Poincare’s theorem. Now,verify that the BF – action is not only invariant under local SL(2,C)−transformations but also under

BIJ = 7→ BIJ + (D ∧ θ)IJ = BIJ + dθIJ + ωI K ∧ θKJ + θIK ∧ ωK J (III.3.2.4)

for some sl(2,C) valued one – form θ and that therefore also B can be gauged to zero locally.Hint:Use the Bianchi identity for Ω.ii) Perform the Legendre transformation and conclude that there are as many first class constraintsas canonical pairs so that again at most a countable number of global degrees of freedom can exist.

One may wonder how it is possible that a theory with less kinematical degrees of freedom has moredynamical (true) degrees of freedom. The answer is that BF – theory has by far more symmetriesthan the Palatini theory, thus when constraining the number of degrees of freedom we are freezingmore symmetries than we deleted degrees of freedom.

10Another solution is BIJ = eI ∧ eJ but this possibility is currently not discussed.

60

Let us now discuss how one formulates the path integral corresponding to the action (III.3.2.3).It is formally given by

KP (Σt1 ,Σt2) =

∫[Dω DB DΦ]eiS

′P [ω,B,Φ] (III.3.2.5)

where Σt1,Σt2 denote suitable boundary conditions specified in more detail below. Suppose we setΦ = 0, then (III.3.2.5) is a path integral for BF – theory and the integral over B results in thefunctional δ−distribution δ[Ω] imposing the flatness of ω. Now flatness of a connection is equivalentto trivial holonomy along contractable loops by the Ambrose – Singer theorem. If one regularizes thepath integral by introducing a triangulation τ of M , then δ[F ] can be written as

∏α δ(ω(α), 1) where

the product is over a generating system of independent, contractible loops in τ and δ(ω(α), 1) denotesthe δ−distribution on SL(2,C) with respect to the Haar measure. Since SL(2,C) is a non-compactgroup, the δ−distribution is a direct integral over irreducible, unitary representations rather than adirect sum as it would be the case for compact groups (Peter&Weyl theorem). Such representationsare infinite dimensional and are labelled by a continuous parameter ρ ∈ R+

0 and a discrete parametern ∈ N+

0 . Thus, one arrives at a triangulated a spin foam model: For a fixed triangulation oneintegrates (sums) over all possible “spins” ρ (n) that label the generating set of loops (equivalently:the faces that they enclose) of that triangulated four manifold. The analogy with the state summodels for TQFT’s is obvious.

Now what one does is a certain jump, whose physical implication is still not understood: Insteadof performing perturbation theory in SI one argues that formally integrating over Φ and thus im-posing the simplicity constraint is equivalent to the restriction of the direct integral that enters theδ−distributions to simple representations, that is, representations for which either n = 0 or ρ = 0.In other words, one says that the triangulated Palatini path integral is the same as the triangulatedBF path integral restricted to simple representations. To motivate this argument, one notices thatupon canonical quantization of BF theory on a triangulated manifold the B field is the momentumconjugate to ω and if one quantizes on a Hilbert space based on sl(2,C) connections using the Haarmeasure (similar as we have done for SU(2) for a fixed graph), its corresponding flux operator BIJ(S)becomes a linear combination of right invariant vector fields RIJ on SL(2,C). Now the simplicityconstraint becomes the condition that the second Casimir operator RIJRKLǫIJKL vanishes. However,on irreducible representations this operator is diagonal with eigenvalues nρ/4. While this is a strongmotivation, it is certainly not sufficient justification for this way of implementing the simplicityconstraint in the path integral because it is not clear how this is related to integrating over Φ.

In any case, if one does this then one arrives at (some version of) the Lorentzian Barrett – Cranemodel [59]. Surprisingly, for a large class of triangulations τ the amplitudes

KτP (Σt1 ,Σt2) := [

∫[Dω DB]eiSBF [ω,B]]|simplereps. (III.3.2.6)

actually converge although one integrates over a non-compact group ! This is a non-trivial result[60]. The path integral is then over all possible representations that label the faces of a spin foamand the boundary conditions keep the representations on the boundary graphs, that is, spin networksfixed (SL(2,C) reduces to the SU(2) on the boundary). This also answers the question of what theboundary conditions should be.

There is still an open issue, namely how one should get rid of the regulator (or triangulation)dependence. Since BF – theory is a topological QFT, the amplitudes are automatically triangulationindependent, however, this is certainly not the case with GR. One possibility is to sum over trian-gulations and a concrete proposal of how to weigh the contributions from different triangulationscomes from the so-called field theory formulation of the theory [61]. Here one reformulates the BF

61

– theory path integral as the path integral for a scalar field on a group manifold which in this caseis a certain power of SL(2,C). The action for that scalar field has a free piece and an interactionpiece and performing the perturbation theory (Feynman graphs !) for that field theory is equiva-lent to the sum over BF – theory amplitudes for all triangulated manifolds with precisely definedweights. This idea can be straightforwardly applied also to our context where the restriction to sim-ple representations is realized by imposing corresponding restrictions (projections) on the scalar field.

Summarizing, spin foam models are a serious attempt to arrive at a covariant formulation of QGRbut many issues are still unsettled, e.g.:1)There is no clean equivalence with the Hamiltonian formulation as we have seen. Without that itis unlcear how to interpret the spin foam model amplitude and whether it has the correct classicallimit. In order to make progress on the issue of the classical limit, model independent techniques forconstructing “causal spin foams” [62] with a built in notion of quantum causality and renormalizationmethods [63], which should allow in principle the derivation of a low energy effective action, havebeen developed.2)The physical correctness of the Barrett – Crane model is unclear. This is emphasized by recentresults within the Euclidean formulation [64] which suggest that the classical limit is far off GR sincethe amplitudes are dominated by spin values close to zero. This was to be expected because in thedefinition of the Barrett – Crane model there is a certain flexibility concerning the choice of themeasure that replaces [Dω DB] at the triangulated level and the result [64] indicates that one mustgain more control on that choice.3)It is not even clear that these models are four – dimensionally covariant: One usually defines thatthe amplitudes for a fixed triangulation are the same for any four - diffeomorphic triangulation.However, recent results [65] show that this natural definition could result nevertheless in anomalies.This problem is again related to the choice of the measure just mentioned.

Thus, substantially more work is required in order to fill in the present gaps but the results al-ready obtained are very promising indeed.

III.4 Quantum Black Holes

III.4.1 Isolated Horizons

Any theory of quantum gravity must face the question whether it can reproduce the celebrated resultdue to Bekenstein and Hawking [66] that a black hole in a spacetime (M, g) should account for aquantum statisical entropy given by

SBH =Ar(H)

4ℓ2P(III.4.1.1)

where H denotes the two-dimensional event horizon of the blak hole. This result was obtained withinthe framework of QFT on Curved SpaceTimes (CST) and should therefore be valid in a semiclassicalregime in which quantum fluctuations of the gravitational field are neglible (large black holes). Themost important question from the point of view of a microscopical theory of quantum gravity is,what are the microscopical degrees of freedom that give rise to that entropy. In particular, how can

62

it be within a quantum field theory with an infinite number of degrees of freedom, that this entropy,presumably a measure for our lack of information of what happens behind the horizon, comes outfinite.

In [67] the authors performed a bold computation: For any surface S and any positive numberA0 they asked the question how many SNW states there are in QGR such that the area operatorseigenvalues lie within the interval [A0 − ℓ2P , A0 + ℓ2P ]. This answer is certainly infinite because aSNW can intersect S in an uncountably infinite number of different positions without changingthe eigenvalues. This divergence can be made less severe by moding out by spatial diffeomorphismswhich we can use to map these different SNW onto each other in the vicinity of the surface. However,since there are still an infinite number of non-spatially difeomorphic states which look the same inthe vicinity of the surface but different away from it, the answer is still divergent. Therefore, onehas to argue that one must not count information off the surface, maybe invoking the Hamiltonianconstraint or using the information that S = H is not an arbitrary surface but actually the horizonof a black hole. Given this assumption, the result of the, actually quite simple counting problemcame rather close to (III.4.1.1) with the correct factor of 1/4.

Thus the task left is to justify the assumptions made and to make the entropy counting water– tight by invoking the information that H is a black hole horizon. The outcome of this analysiscreated a whole industry of its own, known under the name “isolated horizons”, which to large partis a beautiful new chapter in classical GR. In what follows we will focus only on a tiny fraction ofthe framework, mostly concentrating on the ingredients essential for the quantum formulation. Forreviews see [68] which also contain a complete list of references on the more classical aspects of thisprogramme, the pivotal papers concerning the quantum applications are [69].

By definition, an event horizon is the external boundary of the part of M that does not lie inthe past of null future infinity J+ in a Penrose diagramme. From an operational point of view,this definition makes little sense because in order to determine whether a candidate is an eventhorizon, one must know the whole spacetime (M, g) which is never possible by measurements whichare necessarily loacal in spacetime (what looks like an eternal black hole now could capture somedust later and the horizon would change its location). Thus one looks for some local substitute of thenotion of an event horizon which captures the idea that the black hole has come to some equilibriumstate at least for some amount of time. This is roughly what an isolated horizon ∆ is, illustrated infigure 14. More in technical details we have the following.

Definition III.4.1.A part ∆ of the boundary ∂M of a spacetime (M, g) is called an isolated horizon, provided that1) ∆ ≡ R × S2 is a null hypersurface and has zero shear and expansion11.2) The field equations and matter energy conditions hold at ∆.3) g is Lie derived by the null generator l of ∆ at ∆.

The canonical formulation of a field theory on a manifold M with boundary ∆ must involveboundary conditions at ∆ in order that the variation principle be well-defined (the action must befunctionally differentiable). Such boundary conditions usually give birth to boundary degrees offreedom [70] which would normally be absent but now come into being because (part of the) gaugetransformations are forced to become trivial at ∆. In the present situation what happens is that the

11Recall the notions of shear, expansion and twist of a congruence of vector fields in connection with Raychaudhuri’sequation.

63

Figure 14: An isolated horizon ∆ boundary of a piece M (shaded) of spacetime also bounded byspacelike hypersurfaces Σ1,Σ2. Radiation γ may enter or leave M and propagate into the singularitybefore or after the isolated horizon has formed but must not cross ∆. An intersection of a spacelikehypersurface Σ with ∆ is denoted by H which has the topology of a sphere.

boundary term is actually a U(1) Chern-Simons action12

SCS =A0

πβ

W ∧ dW =

R

dt

H

d2yǫIJ [WIWJ +Wt(dW )IJ ] (III.4.1.2)

where W is a U(1) connection one form and H = S2 = Σ ∩ ∆ is a sphere. The relation between thebulk fields Aja, E

aj and the boundary fields WI , I = 1, 2 is given by

X∗HA

j = Wrj and [X∗H(∗E)j]rj = − A0

2πβdW (III.4.1.3)

where XH : H → Σ is the embedding of the boundary H of Σ into Σ and rj is an arbitrary but fixedunit vector in su(2) which is to be preserved under SU(2) gauge transformations at ∆ and thereforereduces SU(2) to U(1). The number A0 is the area of H as measured by g which turns out to be aconstant of the motion as a consequence of the field equations. The existence of rj is a consequenceof definition (III.4.1) and ∗E is the natural metric independent two – form dual to E.

III.4.2 Entropy Counting

One now has to quantize the system. This consists of several steps whose details are complicatedand which we will only sketch in what follows.

i) Kinematical Hilbert SpaceThe bulk and boundary degrees of freedom are independent of each other, therefore we chooseHkin = HΣ

kin ⊗ HHkin where both spaces are of the form L2(A, dµ0) just that the first factor is

for an SU(2) bundle over Σ while the second is for an U(1) bundle over H .

12It was observed first in [71] that general relativity in terms of connection variables and in the presence of boundariesleads to Chern – Simons boundary terms.

64

ii) Quantum Boundary conditionsEquation (III.4.1.3) implies, in particular, that in quantum theory we must have schematically

[ [X∗H(∗E)j]rj ] ⊗ idHH

kin= idHΣ

kin⊗ [− A0

2πβdW ] (III.4.2.1)

Now we have seen in the bulk theory that we have discussed in great detail throughout thisreview, that ∗E is an operator valued distribution which must be smeared by two – surfacesin order to arrive at the well-defined electric fluxes. Since (III.4.2.1) is evaluated at H , thisflux operator will non-trivially act only on SNWF’s Ts which live in the bulk but intersect Hin punctures p ∈ H ∩ γ(s). Now the distributional character of the electric fluxes implies thatthe left hand side of (III.4.2.1) is non-vanishing only at those punctures. Thus the curvatureof W is flat everywhere except for the punctures where it is distributional.

Consider now SNWF’s Ts of the bulk theory and those of the boundary theory T ′c. Then

[X∗H(∗E)j]rj ] acts on Ts like the z−component of the angular momentum operator and will

have distributional eigenvalues proportional to the magnetic quantum numbers me of the edgeswith punctures p = e ∩H and spin je where me ∈ −je,−je + 1, .., je.

iii) Implementation of Quantum Dynamics at ∆

It turns out that X∗H(∗E)j]rj and dW are the generators of residual SU(2) gauge transforma-

tions close to XH(H) and of U(1) on H respectively. Now these residual SU(2) transformationsare frozen to U(1) transformations by rj and the most general situation in order for a stateto be gauge invariant is that these residual SU(2) transformations of the bulk theory and theU(1) transformations of the boundary theory precisely cancel each other. It turns out that thiscancellation condition is precisely given by the quantum boundary condition (III.4.2.1). Thusthe states that solve the Gauss constraint are linear combinations of states of the form Ts⊗ T ′

c

where the boundary data of these states are punctures p ∈ P where p ∈ γ(s) ∩ H , the spinsjp = jep of edges e ∈ E(γ(s)) with ep∩H = p and their magnetic quantum numbers mp = mep .However, due to the specific features of the geometrical quantization of Chern-Simons theories[72] the mp cannot be specified freely, they have to satisfy the constraint

p∈P2mp = 0 mod k, k =

A0

4πℓ2P(III.4.2.2)

where k is called the level of a Chern Simons theory which is constrained to be an integer dueto Weil’s quantization obstruction cocycle criterion of geometric quantization [30] and comesabout as follows: The T ′

c are actually fixed to be Θ−functions of level k labelled by integers apwhich satisfy the gauge invariance condition

2mp = −ap mod k,∑

p

ap = 0 mod k (III.4.2.3)

Next, the spatial diffeomorphism constraint of the bulk theory tells us that the position of thepunctures on H are not important, important is only their number.

Finally, one of the boundary conditions at ∆ implies that the lapse becomes trivial N = 0at H if C(N) is to generate an infinitesimal time reparameterization13. Thus, luckily we can

13This does not mean that the lapse of a classical isolated horizon solution must vanish at S, rather there is a subtledifference between gauge motions and symmetries for field theories with boundaries [70] where in this case symmetriesmap solutions to gauge inequivalent or equivalent ones respectively, if N|H 6= 0 or N|H = 0 respectively.

65

H

Bulk

p_1

p_2

p_3

j_2,m_2

j_3,m_3

j_1,m_1

Figure 15: Punctures, spins, magnetic quantum numbers and entropy counting. Only the relevantboundary data are shown, the bulk information is traced over.

escape the open issues with the Hamiltonian constraint as far as the quantum dynamics at His concerned.

We can now come to the issue of entropy counting. First of all we notice that Ar(H) is a Diracobservable because H is invariant under Diff(H) and N = 0 at H . Given n punctures with spinsjl, l = 1, .., N the area eigenvalue for H is

λ(n,~j) = 8πℓ2Pβn∑

n=1

√jl(jl + 1) (III.4.2.4)

Now the physical Hilbert space is of the form

Hphys = ⊕n,~j,~m,~a=k−2~m HBn,~j,~m

⊗HBHn,~j,~m

⊗HHn,~a (III.4.2.5)

where HBHn,~j,~m

describes bulk degrees of freedom at H corresponding to the black hole (finite dimen-

sional), HBn,~j,~m

describes bulk degrees of freedom away from H and finally HHn,~a describes Chern –

Simons degrees of freedom which are completetly fixed in tems of ~m due to reasons of gauge invariance(III.4.2.3). The situation is illustrated in figure 15. Let δ > 0 and let SA0,δ be the set of eigenstatesψn,~j,~m ∈ HBH

n,~j,~mof the area operator such that the eigenvalue lies in the interval [A0 − δ, A0 + δ] and

NA0,δ their number. Define the density matrix

ρBH = idB ⊗ [1

N A0,δ

ψ∈SA0,δ

|ψ >< ψ|] ⊗ idH (III.4.2.6)

The quantum statistical entropy from this microcanonical ensemble is given by

SBH = −Tr(ρBH ln(ρBH)) = ln(NA0,δ) (III.4.2.7)

Thus we just need to count states and the answer will be finite because the area operator has an areagap.

66

Exercise III.4.1.Estimate NA0,δ from above and below taking into account the constraint (III.4.2.2) and that k is aninteger (purely combinatorical problem !).

The result of the counting problem is that SBH is indeed given by (III.4.1.1) to leading order inA0 (there are logarithmic corrections) for δ ≈ ℓ2P provided that

β =ln(2)

π√

3(III.4.2.8)

Here the numbers ln(2),√

3 comes from the fact that the configurations with lowest spin jl = 1/2make the dominant contribution to the entropy with eigenvalue ∝ βn

√3 ≈ A0 and number of states

given by NA0,δ ≈ 2n that is, two Boolean degrees of freedom per puncture [73]. This provides anexplicit explanation for the origin of the entropy. Now fixing β at the value (III.4.2.8) would makelittle sense would it be different for different types of black hole (that is, in presence of differentmatter, charges, rotation, other hair,..). However, this is not the case !

In summary, the analysis sketched above provides a self-contained derivation of SBH within QGR.The result is highly non-trivial because it was not to be expected from the outset that Loop Quan-tum Gravity, classical GR and Chern Simons theory would interact in such a harmonic way as toprovide the expected result: Chern – Simons theory is very different from QGR and still they havean interface at H . The result applies to astrophysically interesting balck holes of the Schwarzschildtype and does not require supersymmetry. Nevertheless, the calculation still has a semiclassical inputbecause the presence of the isolated horizon is fed in at the classical level already. It would be moresatisfactory to have a quantum definition of an (isolated) horizon but this is a hard task and left forfuture research. Another unsolved problem then is the calculation of the Hawking effect from firstprinciples.

III.5 Semiclassical Analysis

III.5.1 The Complexifier Machinery for Generating Coherent States

Let us first specify what we mean by semiclassical states.

Definition III.5.1.Let be given a phase space M, ., . with preferred Poisson subalgebra O of C∞(M) and a Hilbertspace H, [., .] together with an operator subalgebra O of L(H). The triple M, ., .,O is said tobe a classical limit of the triple H, [., .], O provided that there exists an (over)complete set of statesψmm∈M such that for all O,O′ ∈ O the infinitesimal Ehrenfest property

|< O >m

O(m)− 1| ≪ 1

|< [O, O′] >m

i~O,O′(m)− 1| ≪ 1 (III.5.1.1)

and the small fluctuation property

|< O2 >m

< O >2m

− 1| ≪ 1 (III.5.1.2)

67

holds at generic14 points in M. Here < . >m:=< ψm, .ψm > /||ψm||2 denotes the expectation valuefunctional.

For systems with constraints, strictly speaking, semiclassical states should be physical states, thatis, those that solve the constraints because we are not interested in approximating gauge degrees offreedom but only physical observables. Only then are the predictions (~ corrections to the classicallimit) of the theory reliable. In the present situation with QGR, however, we are more interested inconstructing kinematical semiclassical states for the following reason: As we have shown, the statusof the physical correctness of the Hamiltonian constraint operator C is unsettled. We would thereforelike to test whether it has the correct classical limit. This test is obviously meaningless on stateswhich the Hamiltonian constraint annihilates anyway. For the same reason it also does not makesense to construct semiclassical states which are at least spatially diffeomorphism invariant becausethe Hamiltonian constraint does not leave this space invariant.

The key question then is how to construct semiclassical states. Fortunately, for phase spaceswhich have a cotangent bundle structure as is the case with QGR, a rather general constructionguideline is available [74], the so-called Complexifier Method, which we will now sketch:

Let (M, ., .) be a phase space with (strong) symplectic structure ., . (notice that M is allowedto be infinite dimensional). We will assume that M = T ∗C is a cotangent bundle. Let us then choosea real polarization of M, that is, a real Lagrangean submanifold C which will play the role of ourconfiguration space. Then a loose definition of a complexifier is as follows:

Definition III.5.2. A complexifier is a positive definite function 15 C on M with the dimensionof an action, which is smooth a.e. (with respect to the Liouville measure induced from ., .) andwhose Hamiltonian vector field is everywhere non-vanishing on C. Moreover, for each point q ∈ C thefunction p 7→ Cq(p) = C(q, p) grows stronger than linearly with ||p||q where p is a local momentumcoordinate and ||.||q is a suitable norm on T ∗

q (C).

In the course of our discussion we will motivate all of these requirements.The reason for the name complexifier is that C enables us to generate a complex polarization of

M from C as follows: If we denote by q local coordinates of C (we do not display any discrete orcontinuous labels but we assume that local fields have been properly smeared with test functions)then

z(m) :=

∞∑

n=0

in

n!q, C(n)(m) (III.5.1.3)

define local complex coordinates of M provided we can invert z, z for m := (q, p) where p are thefibre (momentum) coordinates of M. This is granted at least locally by definition III.5.2. Here themultiple Poisson bracket is inductively defined by q, C(0) = q, q, C(n+1) = q, C(n), C andmakes sense due to the required smoothness. What is interesting about (III.5.1.3) is that it impliesthe following bracket structure

z, z = z, z = 0 (III.5.1.4)

while z, z is necessarily non-vanishing. The reason for this is that (III.5.1.3) may be written in themore compact form

z = e−iLχC q = ([ϕtχC]∗q)t=−i (III.5.1.5)

14The set of points where (III.5.1.1), (III.5.1.2) are violated should have small Liouville measure.15For the rest of this section C will denote a complexifier function and not the Hamiltonian constraint.

68

where χC denotes the Hamiltonian vector field of C, L denotes the Lie derivative and ϕtχCis the one

– parameter family of canonical transformations generated by χC . Formula (III.5.1.5) displays thetransformation (III.5.1.3) as the analytic extension to imaginary values of the one parameter familyof diffeomorphisms generated by χC and since the flow generated by Hamiltonian vector fields leavesPoisson brackets invariant, (III.5.1.4) follows from the definition of a Lagrangean submanifold. Thefact that we have continued to the negative imaginary axis rather than the positive one is importantin what follows and has to do with the required positivity of C.

The importance of this observation is that either of z, z are coordinates of a Lagrangean sub-manifold of the complexification MC, i.e. a complex polarization and thus may serve to define aBargmann-Segal representation of the quantum theory (wave functions are holomorphic functions ofz). The diffeomorphism M → CC; m 7→ z(m) shows that we may think of M either as a symplecticmanifold or as a complex manifold (complexification of the configuration space). Indeed, the polar-ization is usually a positive Kahler polarization with respect to the natural ., .-compatible complexstructure on a cotangent bundle defined by local Darboux coordinates, if we choose the complexifierto be a function of p only. These facts make the associated Segal-Bargmann representation especiallyattractive.

We now apply the rules of canonical quantization: a suitable Poisson algebra O of functions Oon M is promoted to an algebra O of operators O on a Hilbert space H subject to the conditionthat Poisson brackets turn into commutators divided by i~ and that reality conditions are reflectedas adjointness relations, that is,

[O, O′] = i~ O,O′ + o(~), O† = ˆO + o(~) (III.5.1.6)

where quantum corrections are allowed (and in principle unavoidable except if we restrict O, say tofunctions linear in momenta). We will assume that the Hilbert space can be represented as a space ofsquare integrable functions on (a distributional extension C of) C with respect to a positive, faithfulprobability measure µ, that is, H = L2(C, dµ) as it is motivated by the real polarization.

The fact that C is positive motivates to quantize it in such a way that it becomes a self-adjoint,positive definite operator. We will assume this to be the case in what follows. Applying then thequantization rules to the functions z in (III.5.1.3) we arrive at

z =

∞∑

n=0

in

n!

[q, C](n)

(i~)n= e−C/~qeC/~ (III.5.1.7)

The appearence of 1/~ in (III.5.1.7) justifies the requirement for C/~ to be dimensionless in definitionIII.5.2. We will call z annihilation operator for reasons that will become obvious in a moment.

Let now q 7→ δq′(q) be the δ-distribution with respect to µ with support at q = q′. (More inmathematical terms, consider the complex probability measure, denoted as δq′dµ, which is definedby

∫δq′dµf = f(q′) for measurable f). Notice that since C is non-negative and necessarily depends

non-trivially on momenta (which will turn into (functional) derivative operators in the quantum

theory), the operator e−C/~ is a smoothening operator. Therefore, although δq′ is certainly not square

integrable, the complex measure (which is probability if C · 1 = 0)

ψq′ := e−C/~δq′ (III.5.1.8)

has a chance to be an element of H. Whether or not it does depends on the details of M, ., ., C.For instance, if C as a function of p at fixed q has flat directions, then the smoothening effect of

69

e−C/~ may be insufficient, so in order to avoid this we required that C is positive definite and notmerely non-negative. If C would be indefinite, then (III.5.1.8) has no chance to make sense as an L2

function.We will see in a moment that (III.5.1.8) qualifies as a candidate coherent state if we are able

to analytically extend (III.5.1.8) to complex values z of q′ where the label z in ψz will play therole of the point in M at which the coherent state is peaked. In order that this is possible (andin order that the extended function is still square integrable), (III.5.1.8) should be entire analytic.Now δq′(q) roughly has an integral kernel of the form ei(k,(q−q

′)) (with some pairing < ., . > betweentangential and cotangential vectors) which is analytic in q′ but the integral over k, after applying

e−C/~, will produce an entire analytic function only if there is a damping factor which decreases fasterthan exponentially. This provides the intuitive explanation for the growth requirement in definitionIII.5.2. Notice that the ψz are not necessarily normalized.

Let us then assume that

q 7→ ψm(q) := [ψq′(q)]q′→z(m) = [e−C/~δq′(q)]q′→z(m) (III.5.1.9)

is an entire L2 function. Then ψm is automatically an eigenfunction of the annihilation operator zwith eigenvalue z since

zψm = [e−C/~qδq′ ]q′→z(m) = [q′e−C/~δq′ ]q′→z(m) = z(m)ψm (III.5.1.10)

where in the second step we used that the delta distribution is a generalized eigenfunction of theoperator q. But to be an eigenfunction of an annihilation operator is one of the accepted definitionsof coherent states !

Next, let us verify that ψm indeed has a chance to be peaked at m. To see this, let us considerthe self-adjoint (modulo domain questions) combinations

x :=z + z†

2, y :=

z − z†

2i(III.5.1.11)

whose classical analogs provide real coordinates for M. Then we have automatically from (III.5.1.10)

< x >m:=< ψm, xψm >

||ψm||2=z(m) + z(m)

2=: x(m) (III.5.1.12)

and similar for y. Equation (III.5.1.12) tells us that the operator z should really correspond to thefunction m 7→ z(m), m ∈ M.

Now we compute by similar methods that

< [δx]2 >m:=< ψm, [x− < x >m]2ψm >

||ψm||2=< [δy]2 >m=

1

2| < [x, y] >m | (III.5.1.13)

so that the ψm are automatically minimal uncertainty states for x, y, moreover the fluctuations areunquenched (equal each other). This is the second motivation for calling the ψm coherent states.Certainly one should not only check that the fluctuations are minimal but also that they are smallas compared to the expectation value, at least at generic points of the phase space, in order that thequantum errors are small.

The infinitesimal Ehrenfest property

< [x, y] >z

i~= x, y(m) +O(~) (III.5.1.14)

70

follows if we have properly implemented the canonical commutation relations and adjointness rela-tions. The size of the correction, however, does not follow from these general considerations butthe minimal uncertainty property makes small corrections plausible. Condition (III.5.1.14) suppliesinformation about how well the symplectic structure is reproduced in the quantum theory.

For the same reason one expects that the peakedness property

|< ψm, ψm′ > |2||ψm||2 ||ψm′ ||2 ≈ χKm(m′) (III.5.1.15)

holds, where Km is a phase cell with center m and Liouville volume ≈√< [δx]2 >m< [δy]2 >m and

χ denotes the characteristic function of a set.Finally one wants coherent states to be overcomplete in order that every state in H can be

expanded in terms of them. This has to be checked on a case by case analysis but the fact that ourcomplexifier coherent states are for real z nothing else than regularized δ distributions which in turnprovide a (generalized) basis makes this property plausible to hold.

Exercise III.5.1.Consider the phase space: M = T ∗R = R2 with standard Poisson brackets q, q = p, p =0, p, q = 1 and configuration space C = R. Consider the complexifier C = p2/(2σ) where σ is adimensionful constant such that C/~ is dimensionfree. Check that it meets all the requirements ofdefinition III.5.2 and perform the coherent state construction displayed above.Hint:Up to a phase, the resulting, normalized coherent states are the usual ones for the harmonic oscillatorwith Hamiltonian H = (p2/m + mω2q2)/2 with σ = mω. Verify that the states ψm are Gaussianpeaked in the configuration representation with width

√~/σ around q = q0 and in the momentum

representation around p = p0 with width√

~σ where m = (p0, q0).

As it has become clear from the discussion, the complexifier method gives a rough guideline,but no algorithm, in order to arrive at a satisfactory family of coherent states, there are things tobe checked on a case by case basis. On the other hand, what is nice is that given only one input,namely the complexifier C, it is possible to arrive at a definite and constructive framework for asemiclassical analysis. It is important to know what the classical limit of C is, otherwise, if we havejust an abstract operator C then the map m 7→ z(m) is unknown and an interprtation of the statesin terms of M is lost.

III.5.2 Application to QGR

Let us now apply these ideas to QGR. Usually the choice of C is stronly motivated by a Hamiltonian,but in QGR we have none. Therefore, at the moment the best we can do is to play with variousproposals for C and to explore the properties of the resulting states. For the simplest choice of C[75] those properties have been worked out more or less completely and we will briefly describe thembelow.

The operator C is defined by its action on cylindrical functions f = p∗γfγ by

C

~f = −p∗γ [

1

2[

e∈E(γ)

l(e)[Rje/2]2] fγ ] =: p∗γ [Cγfγ] (III.5.2.1)

where the positive numbers l(e) satisfy l(e e′) = l(e) + l(e′) and l(e−1) = l(e) and Rje are the usual

right invariant vector fields.

71

Exercise III.5.2.Recall the definition of the maps pγ′γ for γ ≺ γ′ from section II.1.2 and check that the Cγ are

consistently defined, that is, Cγ′ p∗γ′γ = p∗γ′γ Cγ.

This choice is in analogy to the harmonic oscillator where the quantum complexifier is essentiallythe Laplacian −(d/dx)2. The classical limit of (III.5.2.1) depends in detail on the function l whichis analogous to the parameter ~/σ for the case of the harmonic oscillator. For instance [74], one canchoose a) three families of foliations s 7→ HI

s , I = 1, 2, 3 of σ by two dimensional surfaces HIs such

that there is a bijection (s1, s2, s3) 7→ x(~s) := [∩IHIsI ] ∈ σ and b) a partition P I

s of the HIs into small

surfaces S and define

C =1

2a2κ

R

ds3∑

I=1

S∈P Is

[Ar(S)]2 (III.5.2.2)

where Ar(S) is again the area functional and a is a dimensionful constant of dimension cm1. The

function l for this example is then roughly16 l(e) = (βℓP )2

a2

∫ds

∑I

∑S∈P I

sχS(e) where χS(e) = 1 if

S ∩ e 6= ∅ and vanishes otherwise.The δ−distribution with respect to the measure µ0 can be written as the sum over all SNW’s

(exercise !)

δA′(A) =∑

s

Ts(A′)Ts(A) (III.5.2.3)

with resulting coherent states

ψAC(A) =∑

s

e−12

∑e∈E(γ(s)) l(e)je(je+1)Ts(A

C)Ts(A) (III.5.2.4)

where the SL(2,C) connection AC is defined by

AC[A,E] =

∞∑

n=0

in

n!(A,C(n))[A,E] (III.5.2.5)

Thus we see that in this case the symplectic manifold given as the cotangential bundle M = T ∗Aover the space of SU(2) connections is also naturally given as the complex manifold AC of SL(2,C)connections. From the general discussion above it now follows that the classical interpretation of theannihilation operators

AC(e) := e−C/~A(e)eC/~ (III.5.2.6)

is simply the holonomy of the complex connection AC(e).In order to study the semiclassical properties of these states we consider their cut-offs ψAC,γ for

each graph γ defined on cylindrical functions f = p∗γfγ by

< ψAC , f >kin=:< ψAC,γ, f >kin (III.5.2.7)

Now, (if we work at the non-gauge invariant level,) one can check that

ψAC,γ(A) =∏

e∈E(γ(s)

ψl(e)

AC(e)(A(e)) (III.5.2.8)

16This formula gets exact in the limit of infinitely fine partition, at finite coarseness, it is an approximation to theexact, more complicated formula.

72

where for any g ∈ SL(2,C), h ∈ SU(2) we have defined

ψtg(h) :=∑

j

(2j + 1)e−tj(j+1)/2χj(gh−1) (III.5.2.9)

Exercise III.5.3.Verify, using the Peter&Weyl theorem, that for g ∈ SU(2) we have ψ0

g(h) = δg(h), the δ−distributionwith respect to L2(SU(2), dµH). Conclude that (III.5.2.9) is just the analytic extension of the heatkernel e−t∆/2 where ∆ is the Laplacian on SU(2). Thus the states (III.5.2.9) are in complete analogywith those for the harmonic oscillator, just that R was replaced by SU(2) and the complexificationC of R by the complexification SL(2,C) of SU(2). In this form, coherent states on compact gaugegroups were originally proposed by Hall [76].

The analysis of the semiclassical properties of the states ψAC,γ on Hkin can therefore be reducedto that of the states ψtg on L2(SU(2), dµH). We state here without proof that the following propertiescould be proved [75]: I) Overcompleteness, II) expectation value property, III) Ehrenfest property,IV) peakedness in phase space, V) annihilation operator eigenstate property, VI) minimal uncertaintyproperty and VII) small fluctuation property. Thus, these states have many of the desired propertiesthat one requires from coherent states.

In the following graphic we display as an example the peakedness properties of the analog of(III.5.2.9) for the simpler case of the gauge group U(1), the case of SU(2) is similar but requiresmore plots because of the higher dimensionality of SU(2). Thus g0 = eph0 ∈ U(1)C = C − 0, p ∈R, h0 ∈ U(1) and u ∈ U(1) where we parameterize u = eiφ, φ ∈ [−π, π). Similarly, g = ep1u, p1 ∈ R.We consider in figure 16 the peakedness in the configuration representation given by the probabilityamplitude

u = eiφ 7→ jtg0(u) = |ψtg0(u)|2/||ψtg0||2 (III.5.2.10)

at h0 = 1, p ∈ [−5, 5]. In figure 17 the phase space peakedness expressed by the overlap function

g = ep1u 7→ it(g, g0) =| < ψtg, ψ

tg0 > |2

||ψtg||2 ||ψtg0||2(III.5.2.11)

is shown at fixed p = 0, h0 = 1 for p ∈ [−5, 5], u ∈ U(1). We have made use of the fact (exercise!) that ψg0(u) and < ψg, ψg0 > respectively depend only on the combinations g0u

−1 = ep0hu−1 andgg0 = ep+p1u−1h0 respectively. Therefore, peakedness at u = h0 or g = g0 = eph0 respectively forany h0 is equivalent to peakedness at u = 1 or at g = ep0 respectively for h0 = 1. Both plots arefor the value t = 0.001 and one clearly sees the peak width of

√t ≈ 0.03 when resolving those plots

around the peak as in figure 18, which has a close to Gaussian shape just like the harmonic oscillatorcoherent states have. As a first modest application, these states have been used in order to analyzehow one would obtain, at least in principle, the QFT’s on CST’s (Curved SpaceTime) limit from fullQGR in [49]. In particular, it was possible to perform a detailed calculation concerning the existenceof Poincare invariance violating dispersion relations of photon propagation within QGR which werediscussed earlier at a more phenomenological level in the pioneering papers [77]: The idea is that themetric field is a collection of quantum operators which are not mutually commuting. Therefore itshould be impossible to construct a state which is peaked on, say the Minkowski metric, and whichis a simultaneous eigenstate of all the metric operator components, in other words, there should beno such thing as a Poincar/’e invariant state in full QGR17, already because such an object should

17This seems to contradict the fact that we are even interested in four dimensionally diffeomorphism invariant states

73

Figure 16: Probability amplitude u 7→ jtg0(u) at p ∈ [−5, 5], h0 = 1.

Figure 17: Overlap function g 7→ itg0(g) at p = 0, h0 = 1 for p1 ∈ [−5, 5], u ∈ U(1).

74

Figure 18: Resolution of a neighbourhood of the peak of the function g 7→ itg0(g) at p = 0, h0 = 1.

be highly background dependent. The best one can construct is a coherent state peaked on theMinkowski metric. The small fluctuations that are encoded in that state influence the propagationof matter and these tiny disturbances could accumulate to measurable sizes in so-called γ−ray burstexperiments [78] where one measures the time delay of photons of higher energy as compared tothose of lower energy as they travel over cosmological distances as a result of the energy dependenceof the speed of light. If such an effect exists then it is a non-perturbative one because perturbativelydefined QFT’s on Minkowski space are by construction Poincare invariant (recall e.g. the Wightmanaxioms from section I.1).

These are certainly only first moderate steps. The development of the semiclassical analysis forQGR is still in its very beginning and there are many interesting and new mathematical and phys-ical issues that have to be settled before one can seriously attack the proof that, for instance, theHamiltonian constraint of section III.1 has the correct classcical limit or that full QGR reduces toclassical GR plus the standard model in the low energy regime.

III.6 Gravitons

III.6.1 The Isomorphism

The reader with a strong background in ordinary QFT and/or string theory will have wonderedthroughout these lectures where in QGR the graviton, which plays such a prominent role in the per-turbative, background dependent approaches to quantum gravity, resides. In fact, if one understandsthe graviton, as usually, as an excitation of the quantum metric around Minkowski space, then thereis a clear connection with the semiclassical analysis of the previous section: One should constructa suitable coherent state which is peaked on the gauge invariant phase space point characterizing

and the fact that the Poincare group should be a tiny subgroup thereof. However, this is not the case because werequire the states only to be invariant under diffeomorphisms which are pure gauge and those have to die off at spatialinfinity. Poincare transformations are therefore not gauge transformation but symmetries and what we are saying isthat there are no Poincare symmetric, diffeomorphism gauge invariant states.

75

Minkowski space and identify suitable excitations thereof as gravitons. It is clear that at the momentsuch graviton states from full QGR cannot be constructed, because we would need first to solve theHamiltonian constraint.

However, one can arrive at an approximate notion of gravitons through the quantization of lin-earized gravity: Linearized gravity is nothing else than the expansion of the full GR action aroundthe gauge variant initial data (E0)aj = δaj , (A0)ja = 0 to second order in E − E0, A which resultsin a free, classical field theory with constraints. In fact, the usual notion of gravitons is preciselythe ordinary Fock space quantization of that classical, free field theory [79]. In order to see whetherQGR can possibly accomodate these graviton states, Varadarajan in a beautiful series of papers [80]has carried out a polymer like quantization of that free field theory on a Hilbert space Hkin whichis in complete analogy to that for full QGR, the only difference being that the gauge group SU(2)is replaced by the gauge group U(1)3. While there are certainly large differences between the highlyinteracting QGR theory and linearized gravity, one should at least be able to gain some insight intothe the answer to the question, how a Hilbert space in which the excitations are one dimensional canpossibly describe the Fock space excitations (which are three dimensional).

The problem of describing gravitons within linearized gravity by polymer like excitations is math-ematically equivalent to the simpler problem of describing the photons of the ordinary Fock Hilbertspace HF of Maxwell theory by polymer like excitations within a Hilbert space HP = L2(A, dµ)where A is again a space of generalized U(1) connections with some measure µ thereon. Thus, wedecribe the latter problem in some detail since it requires less space and has the same educationalvalue.

The crucial observation is the following isomorphism I between two different Poisson subalgebrasof the Poisson algebra on the phase space M of Maxwell theory coordinatized by a canonical pair(E,A) defined by a U(1) connection A and a conjugate electric field E: Consider a one-parameterfamily of test functions of rapid decrease which are regularizations of the δ−distribution, for instance

fr(x, y) =e−

||x−y||2

2r2

(√

2πr)3(III.6.1.1)

where we have made use of the Euclidean spatial background metric. Given a path p ∈ P we denoteits distributional form factor by

Xap (x) :=

∫ 1

0

dt pa(t)δ(x, p(t)) (III.6.1.2)

The smeared form factor is defined by

Xap,r(x) :=

∫d3yfr(x, y)X

ap (y) =

∫ 1

0

dt pa(t)fr(x, p(t)) (III.6.1.3)

which is evidently a test function of rapid decrease. Notice that a U(1) holonomy maybe written as

A(p) := ei∫d3xXa

p (x)Aa(x) (III.6.1.4)

and we can define a smeared holonomy by

Ar(p) := ei∫d3xXa

p,r(x)Aa(x) (III.6.1.5)

76

Likewise we may define smeared electric fields as

Ear (x) :=

∫d3yfr(x, y)E

a(y) (III.6.1.6)

If we denote by q the electric charge (notice that in our notation α = ~q2 is the fine structureconstant), then we obtain the following Poisson subalgebras: On the one hand we have smearedholonomies but unsmeared electric fields with

Ar(p), Ar(p′) = Ea(x), Eb(y) = 0, Ea(x), Ar(p) = iq2Xap,r(x)Ar(p) (III.6.1.7)

and on the other hand we have unsmeared holonomies but smeared electric fields with

A(p), A(p′) = Ear (x), E

br(y = 0, Ea

r (x), A(p) = iq2Xap,r(x)hp (III.6.1.8)

Thus the two Poisson algebras are isomorphic and also the ∗ relations are isomorphic, bothEa(x), Ear (x)

are real valued while both A(p), Ar(p) are U(1) valued. Thus, as abstract ∗− Poisson algebras thesetwo algebras are indistinguishable and we may ask if we can find different representations of it. Evenbetter, notice that Ar(p)Ar(p

′) = Ar(p p′), Ar(p)−1 = Ar(p−1) so the smeared holonomy algebra is

also isomorphic to the unsmeared one. Hence there is an algebra ∗−isomorphism I defined on thegenerators by Ir(hp) = hp,r, Ir(Er) = E. One must also show that the Ar(p) are still algebaricallyindependent as are the A(p) [80].

III.6.2 Induced Fock Representation With Polymer – Excitations

Now we know that the unsmeared holonmy algebra is well represented on the Hilbert space Hkin =L2(A, dµ0) while the smeared holonomy algebra is well represented on the Fock Hilbert space HF =L2(S ′, dµF ) where S ′ denotes the space of divergence free, tempered distributions and µF is theMaxwell-Fock measure. These measures are completely characterized by their generating functional

ωF (Ar(p)) := µF (Ar(p)) = e−14α

∫d3xXa

p,r(x)√−∆

−1Xb

p,rδab (III.6.2.1)

since finite linear combinations of the hp,r are dense in HF [80]. Here ∆ = δab∂a∂b denotes theLaplacian and we have taken a loop p rather than an open path so that Xp,r is transversal. Alsounsmeared electric fields are represented through the Fock state ωF by

ωF (Ar(p)Ea(x)Ar(p

′)) = −α2

[Xap,r(x) −Xa

p′,r(x)]ωF (hpp′,r) (III.6.2.2)

and any other expectation value follows from these and the commutation relations.Since ωF defines a positive linear functional we may define a new representation of the algebra

A(p), Ear by

ωr(A(p)) := ωF (Ar(p)) and ωr(A(p)Ear (x)A(p′)) := ωF (Ar(p)E

a(x)Ar(p′)) (III.6.2.3)

called the r−Fock representation. In other words, we have ωr = ωF Ir.Since ωr is a positive linear functional on C(A) by construction there exists is a measure µr on A

that represents ωr in the sense of the Riesz representation theorem (recall II.1.2.12). In [81] Velhinhoshowed that the one – parameter family of measures µr are expectedly mutually singular with respectto each other and with respect to the uniform measure µ0 (that is, the support of one measure is ameasure zero set with respect to the other and vice versa).

77

Result 1: There is a unitary transformation between any of the Hilbert spaces Hr and their im-ages under Ir in the usual Fock space HF . Since finite linear combinations of the Ar(p) for fixedr are still dense in HF [80], there exists indeed a polymer like description of the usual n-photon states.

Recall that the Fock vacuum ΩF is defined to be the zero eigenvalue coherent state, that is, itis annihilated by the annihilation operators

a(f) :=1√2α

∫d3xfa[

4√−∆Aa − i(

4√−∆)−1Ea] (III.6.2.4)

where fa is any transversal smearing field. We then have in fact that ωF (.) =< ΩF , .ΩF >HF. (For

readers familiar with C∗−algebras this means that ΩF is the cyclic vector that is determined byωF through the GNS construction.) The idea is now the following: From (III.6.2.3) we see thatwe can easily answer any question in the r−Fock representation which has a preimage in the Fockrepresentation, we just have to replace everywhere Ar(p), E

a(x) by A(p), Ear (x). Since in the r−Fock

representations only exponentials of connections are defined, we should exponentiate the annihilationoperators and select the Fock vacuum through the condition

eia(f)ΩF = ΩF (III.6.2.5)

In particular, choosing f =√

2α( 4√−∆)−1Xp,r for some loop p we get

e∫d3xXa

p,r[iAa+(√−∆)−1Ea]ΩF = ΩF (III.6.2.6)

Using the commutation relations and the Baker – Campell – Hausdorff formula one can write(III.6.2.6) in terms of Ar(p) and the exponential of the electric field appearing in (III.6.2.6) times anumerical factor. The resulting expression can then be translated into the r−Fock representation.Denoting the translated expression by I−1

r (eia(f)) we now ask the question, whether there existsa state Ωr ∈ Hkin = L2(A, dµ0) such that I−1

r (eia(f))Ωr = Ωr. Remarkably, expanding Ωr into thecharge network basis introduced in section (III.1) one finds a (up to a multiplicative constant) uniquesolution given by

Ωr =∑

c

e−α

2

∑e,e′∈E(γ(c))G

re,e′

ne(c)ne′ (c)Tc (III.6.2.7)

where c = (γ(c), ne(c)e∈E(γ(c))) denotes a charge network (the U(1) analogue of a spin network)and

Gre,e′ =

∫d3xXa

e,r

√−∆

−1Xbe′,rδ

Tab (III.6.2.8)

where δTab = δab − ∂a∆−1∂b denotes the transverse projector.

Exercise III.6.1.Fill in the gaps that lead from (III.6.2.6) to (III.6.2.8).

Let us discuss this result. First of all, (III.6.2.7) is not normalizable with respect to the innerproduct on Hkin and neither are the images of n−photon states or coherent states from HF . Thisseems to indicate that the space Hkin does not play any role for physically interesting states. However,in [74] it was shown that this is not the case: It turns out, that, given a suitable regularization, thatone can indeed obtain the expectation values such as ωr(A(p)) from the formal expression

ωr(A(p)) :=< Ωr, A(p)Ωr >

||Ωr||2(III.6.2.9)

78

where both numerator and denominator are infinite but the fraction is finite.

Result 2: The polymer images of photon states can be obtained as certain limits of states fromHkin which therefore is a valid starting point in order to obtain physically interesting representations.

Moreover, as can be expected from the similarity between the formulas (III.6.2.8) and (III.5.2.4)(for AC = 0 corresponding to vacuum E = A = 0 in the present case), the states Ωr also arise froma complexifier, given in this case by

C =1

2q2

R3

d3x[Ear

√−∆

−1Ebr ]δab (III.6.2.10)

Result 3: The complexifier framework is also able to derive images of n−photon states and usualFock coherent states from the universal input of a complexifier.

We conclude that at least for the linearized theory the question posed at the beginning of thissection could be answered affirmatively: There is indeed a precise framework available for how toaccomodate graviton states into the framework of loop quantum gravity. This is a promising resultand should have an analog in the full theory.

79

Part IV

Selection of Open Research Problems

80

Let us summarize the most important open research problems that have come up during thediscussion in these lectures.

i) Hamiltonian Constraint and Semiclassical StatesThe unsettled correctness of the quantum dynamics is the major roadblock to completing thequantization programme of QGR. In order to make progress a better understanding of thekinematical semiclassical sector of the theory is necessary.

ii) Physical Inner ProductEven if we had the correct Hamiltonian constraint and the complete space of solutions, at themoment there is no really good idea available of how to construct a corresponding physicalinner product because the constraint algebra is not a Lie algebra but an open algebra in theBRST sense so that techniques from rigged Hilbert spaces are not available. A framework forsuch open algebras must be developed so that an inner product can be constructed at least inprinciple.

iii) Dirac ObservablesNot even in classical general relativity do we know enough Dirac observables. For QGR theyare mandatory for instance in order to select an inner product by adjointness conditions andin order to arrive at an interpretation of the final theory. A framework of how to define Diracobservables, at least in principle, even at the classical level, would be an extremely importantcontribution.

iv) Covariant FormulationThe connection between the Hamiltonian and the Spin Foam formulation is poorly understood.Without such a connection e.g. a proof of covariance of the canonical formulation on the onehand and a proof for the correct classical limit of the spin foam formulation on the other cannotbe obtained using the respective other formulation. One should prove a rigorous Feynman –Kac like formula that allows to switch between these complementary descriptions.

v) QFT on CST’s and Hawking Effect from First PrinciplesThe low energy limit of the theory in connection with the the construction of semiclassicalstates must be better understood. Once this is done, fundamental issues such as whetherthe Hawking effect is merely an artefact of an invalid description by QFT’s on CST’s while aquantum theory of gravity should be used or whether it is a robust result can be answered.Similar remarks apply to the information paradoxon associated with black holes etc.

vi) Combinatorial Formulation of the TheoryThe description of a theory in terms of smooth and even analytic structures curves, surfacesetc. at all scales in which the spectra of geometrical operators are discrete at Planck scales isawkward and cannot be the most adequate language. There should be a purely combinatoricalformulation in which notions such as topology, differential structure etc. can only have asemiclassical meaning.

vii) Avoidance of Classical and UV SingularitiesThat certain classical singularities are absent in loop quantum cosmology and that certainoperators come out finite in the full theory while in the usual perturbative formulation theywould suffer from UV singularities are promising results, but they must be better understood.If one could make contact with perturbative formulations and pin – point exactly why in QGR

81

the usual perturbative UV singularities are absent then the theory would gain a lot morerespect in other communities of high energy physicists. There must be some analog of therenormalization group and the running of coupling constants that one usually finds in QFT’sand CST’s. Similar remarks apply to the generalization of the loop quantum cosmology resultto the full theory.

viii) Contact with String (M) – TheoryIf there is any valid perturbative description of quantum gravity then it is almost certainlystring theory. It is conceivable that both string theory and loop quantum gravity are comple-mentary descriptions but by themselves incomplete and that only a fusion of both can reach thestatus of a fundamental theory. To explore these possibilities, Smolin has launched an ambi-tious programme [82] which to our mind so far did not raise the interest that it deserves18. Thecontact arises through Chern – Simons theory which is part of both Loop Quantum Gravityand M – Theory [83] (when considered as the high energy limit of 11 dimensional Supergravity).Another obvious starting point is the definition of M – Theory as the quantum supermembranein 11 dimensions [84], a theory that could be obtained as the quantization of the classicalsupermembrane by our non-perturbative methods. Finally, a maybe even more obvious con-nection could be found through the so-called Pohlmeyer String [85] which appears to be amethod to quantize the string non-perturbatively, without supersymmetry, anomalies or extradimensions, by working directly at the level of Dirac observables which are indeed possible toconstruct explicitly in this case.

We hope to have convinced the reader that Loop Quantum Gravity is an active and lively approachto a quantum theory of gravity which has produced already many non-trivial results and will continueto do so in the future. There are still a huge number of hard but fascinating problems to be solved ofwhich the above list is at most the tip of an iceberg. If at least a tiny fraction of the readers woulddecide to dive into this challenging area and help in this endeavour, then these lectures would havebeen successful.

Acknowledgements

We thank the Heraeus – Stiftung and the organizers, Dominico Giulini, Claus Kiefer and ClausLammerzahl, for making this wonderful and successful meeting possible and the participants forcreating a stimulating atmosphere through long and deep discussions, very often until early in themorning in the “Burgerkeller”.

18That we did not devote a section to this topic in this review is due to the fact that we would need to include anintroduction to M – Theory into these lectures which would require too much space. The interested reader is referredto the literature cited.

82

Bibliography

[1] T. Thiemann, “Introduction to Modern Canonical Quantum General Relativity”, gr-qc/0110034[2] C. Rovelli, “Loop Quantum Gravity”, Review written for electronic journal “Living Reviews”,

gr-qc/9710008; “Strings, Loops and Others : a critical Survey of the present Approaches toQuantum Gravity”, plenary lecture given at 15th Intl. Conf. on Gen. Rel. and Gravitation(GR15), Pune, India, Dec 16-21, 1997, gr-qc/9803024; “Notes for a Brief History of QuantumGravity”, gr-qc/0006061M. Gaul, C. Rovelli, “Loop Quantum Gravity and the Meaning of Diffeomorphism Invariance”,Lect.Notes Phys.541:277-324,2000, [gr-qc/9910079]G. Horowitz, “Quantum Gravity at the Turn of the Millenium”, gr-qc/0011089S. Carlip, “Quantum Gravity: A Progress Report”, Rept.Prog.Phys.64:885,2001 [gr-qc/0108040]A. Ashtekar, “Quantum Mechanics of Geometry”, gr-qc/9901023

[3] E. E. Flanagan, R. M. Wald, “Does Backreaction enforce the averaged null energy condition insemiclassical gravity ?”, Phys. Rev. D54 (1996) 6233-6283, [gr-qc/9602052]

[4] R. M. Wald, “General Relativity”, The University of Chicago Press, Chicago, 1989[5] W. Junker, “Hadamard States, Adiabatic Vacua and the Construction of Physical States for

Scalar Quantum Fields on Curved Spacetime”, Rev. Math. Phys. 8 (1996) 1091 – 1159[6] M.H. Goroff, A. Sagnotti, Phys. Lett. B160 (1985) 81 M.H. Goroff, A. Sagnotti, Nucl. Phys.

B266 (1986) 709[7] S. Deser, “Non-Renormalizability of (Last Hope) D = 11 Supergravity with a Survey of Diver-

gences in Quantum Gravities”, hep-th/9905017[8] O. Lauscher, M. Reuter, “Towards Nonperturbative Renormalizability of Quantum Einstein

Gravity”, Int. J. Mod. Phys. A17 993 (2002), [hep-th/0112089]; “Is Quantum Gravity Nonper-turbatively Renormalizable ?”, Class. Quant. Grav. 19 483 (2002), [hep-th/0110021]

[9] J. Polchinsky, “String Theory”, vol. 1 and 2, Cambridge University Press, Cambridge, 1998[10] D. Marolf, C. Rovelli, “Relativistic Quantum Measurement”, Phys.Rev. D66 (2002) 023510,

[gr-qc/0203056][11] J. B. Hartle, S. W. Hawking, Phys. Rev. D28 (1983) 2960[12] J. Ambjorn, J. Jurkiewicz, R. Loll, “Lorentzian and Euclidean Quantum Gravity: Analytical

and Numerical Results”, hep-th/0001124; “A Non-Perturbative Lorentzian Path Integral forGravity”, Phys. Rev. Lett. 85 (2000) 924, [hep-th/0002050]J. Ambjorn, A. Dasgupta, J. Jurkiewicz, R. Loll, “A Lorentzian Cure For Euclidean Troubles”,Nucl. Phys. Proc. Suppl. 106 (2002) 977-979, [hep-th/0201104]R. Loll, A. Dasgupta, “A Proper Time Cure for the Conformal Sickness in Quantum Gravity”,Nucl. Phys. B606 (2001) 357-379, [hep-th/0103186]

[13] M. J. Gotay, J. Isenberg, J. E. Marsden, “Momentum Maps and Classical Relativistic Fields.Part 1: Covariant Field Theory”, (with the collaboration of Richard Montgomery, Jedrzej Sni-atycki and Philip B. Yasskin) [physics/9801019]H.A. Kastrup, “Canonical Theories of Dynamical Systems in Physics” Phys. Rept. 101 (1983)

83

1I. V. Kanatchikov, “Canonical Structure of Classical Field Theory in the Polymomentum PhaseSpace”, Rept. Math. Phys. 41 (1998) 49-90, [hep-th/9709229]; “On the Field Theoretic Gener-alizations of a Poisson Algebra”, Rept. Math. Phys. 40 (1997) 225, [hep-th/9710069]C. Rovelli, “A Note on the Foundation of Relativistic Mechanics” [gr-qc/0111037]; “. 2. Co-variant Hamiltonian General Relativity”, [gr-qc/0202079]; “Covariant Hamiltonian Formalismfor Field Theory: Symplectic Structure and Hamilton-Jacobi Equation on the Space G”, [gr-qc/0207043]

[14] A. Ashtekar, L. Bombelli, O. Reula, “The Covariant Phase Space of Asymptotically Flat Grav-itational Fields”, in: “Analysis, Geometry and Mechanics: 200 Years After Lagrange”, Ed. byM. Francaviglia, D. Holm, North-Holland, Amsterdam, 1991

[15] R.E. Peierls, Proc. R. Soc. Lond. A 214 (1952) 143B. S. DeWitt, “Dynamical Theory of Groups and Fields”, Gordon & Breach, New York, 1965

[16] C. J. Isham, N. Linden, “Quantum Temporal Logic and Decoherence Functionals in the His-tories Approach to Generalized Quantum Theory”, J. Math. Phys. 35 (1994) 5452-5476 [gr-qc/9405029]; “Continuous Histories and the History Group in generalized Quantum Theory”, J.Math. Phys. 36 (1995) 5392-5408, [gr-qc/9503063]C.J. Isham, N. Linden, K. Savvidou, S. Schreckenberg, “Continuous Time and Consistent His-tories” J. Math. Phys. 39 (1998) 1818-1834, [quant-ph/9711031]

[17] R. B. Griffiths, J. Stat. Phys. 36 (1984) 219; Found. Phys. 23 1601R. Omnes, J. Stat. Phys. 53 (1988) 893; ibid 53 (1988) 933; ibid 53 (1988) 957; ibid 57 (1989)357; Rev. Mod. Phys. 64 (1992) 339M. Gell-Mann , J. B. Hartle, “Classical Equations for Quantum Systems”, Phys. Rev. D47(1993) 3345-3382,[gr-qc/9210010]; “ Equivalent Sets of Histories and Multiple QuasiclassicalDomains”, [gr-qc/9404013]; “Strong Decoherence”, [gr-qc/9509054]J. B. Hartle, “Space-Time Quantum Mechanics and the Quantum Mechanics of Space-Time”,Les Houches Sum.Sch.1992:0285-480, [gr-qc/9304006]F. Dowker, A. Kent, “Properties of Consistent Histories” Phys.Rev.Lett. 75 (1995) 3038-3041,[gr-qc/9409037]; “On the Consistent Histories Approach to Quantum Mechanics” J. Statist.Phys. 82 (1996) 1575-1646, [gr-qc/9412067]

[18] I. V. Kanatchikov, “Precanonical Perspective in Quantum Gravity”, Nucl. Phys. Proc. Suppl.88 (2000) 326-330, [gr-qc/0004066]; “Precanonical Quantization and the Schroedinger WaveFunctional”, Phys. Lett. A283 (2001) 25-36 [hep-th/0012084]; “Precanonical Quantum Gravity:Quantization without the Spacetime Decomposition”, Int. J. Theor. Phys. 40 (2001) 1121-1149,[gr-qc/0012074]

[19] C.J. Isham, K.N. Savvidou, “Quantizing the Foliation in History Quantum Field Theory”,[quant-ph/0110161]

[20] P. A. M. Dirac, “Lectures on Quantum Mechanics”, Belfer Graduate School of Science, YeshivaUniversity Press, New York, 1964

[21] D. M. Gitman, I. V. Tyutin, “Quantization of Fields with Constraints”, Springer-Verlag, Berlin,1990M. Henneaux, C. Teitelboim, “Quantization of Gauge Systems” Princeton University Press,Princeton, 1992

[22] P. G. Bergmann, A. Komar, “The Phase Space Formulation of General Relativity and Ap-proaches Towards its Canonical Quantization”, Gen. Rel. Grav., 1 (1981) 227-254A. Komar, “General Relativistic Observables via Hamilton Jacobi Functionals”, Phys. Rev. D4(1971) 923-927; “Commutator Algebra of General Relativistic Observables”, Phys. Rev. D9

84

(1974) 885-888; “Generalized Constraint Structure for Gravitation Theory”, Phys. Rev. D27(1983) 2277-2281; “Consistent Factor Ordering of General Relativistic Constraints”, Phys. Rev.D20 (1979) 830-833P. G. Bergmann, A. Komar, “The Coordinate Group Symmetries of General Relativity”, Int. J.Theor. Phys. 5 (1972) 15

[23] S. A. Hojman, K. Kuchar, C. Teitelboim, “Geometrodynamics Regained”, Annals Phys. 96(1976) 88-135

[24] C.G. Torre, I.M. Anderson, “Symmetries of the Einstein Equations” Phys. Rev. Lett. 70 (1993)3525-3529, [gr-qc/9302033]; “Classification of Generalized Symmetries for the Vacuum EinsteinEquations”, Commun. Math. Phys. 176 (1996) 479-539, [gr-qc/9404030]

[25] A. Ashtekar, Phys. Rev. Lett. 57 (1986) 2244, Phys. Rev. D36 (1987) 1587[26] G. Immirzi, Nucl. Phys. Proc. Suppl. 57 (1997) 65[27] F. Barbero, Phys. Rev. D51 (1995) 5507; Phys. Rev. D51 (1995) 5498[28] M. Nakahara, “Geometry, Topology and Physics”, Institute of Physics Publishing, Bristol, 1998[29] D. Giulini, D. Marolf, “On the Generality of Refined Algebraic Quantization”, Class. Quant.

Grav. 16 (1999) 2479-2488 [gr-qc/9812024][30] N. M. J. Woodhouse, “Geometric Quantization”, 2nd. edition, Clarendon Press, Oxford, 1991[31] R. Gambini, A. Trias, Phys. Rev. D22 (1980) 1380

C. Di Bartolo, F. Nori, R. Gambini, A. Trias, Lett. Nuov. Cim. 38 (1983) 497R. Gambini, A. Trias, Nucl. Phys. B278 (1986) 436

[32] M. F. Atiyah, “Topological Quantum Field Theories”, Publ. Math. IHES 68 (1989) 175-186M. F. Atiyah, “The Geometry of Physics and Knots”, Cambridge University Press, Cambridge1990

[33] T. Jacobson, L. Smolin, “Nonperturbative Quantum Geometries” Nucl. Phys. B299 (1988) 295,C. Rovelli, L. Smolin, “Loop Space Representation of Quantum General Relativity”, Nucl. Phys.B331, 80 (1990)

[34] R. Giles, Phys. Rev. D8 (1981) 2160[35] H. Sahlmann, “When Do Measures on the Space of Connections Support the Triad Operators

of Loop Quantum Gravity”, [gr-qc/0207112]; “Some Comments on the Representation Theoryof the Algebra Underlying Loop Quantum Gravity”, [gr-qc/0207111]H. Sahlmann, T. Thiemann, “Representation Theory of Diffeomorphism Invariant Gauge The-ories”, in preparation

[36] A. Ashtekar and J. Lewandowski, J. Math. Phys. 36, 2170 (1995).[37] J. Velhinho, “A Groupoid Approach to Spaces of Generalized Connections”, hep-th/0011200[38] A. Ashtekar, C. J. Isham, Class. Quantum Grav. 9 (1992) 1433[39] A. Ashtekar and J. Lewandowski, “Representation theory of analytic holonomy C⋆ algebras”,

in Knots and quantum gravity, J. Baez (ed), Oxford University Press, Oxford 1994[40] C. Rovelli, L. Smolin, “Spin Networks and Quantum Gravity”, Phys. Rev. D53 (1995) 5743

J. Baez, “Spin Networks in Non-Perturbative Quantum Gravity”, in : “The Interface of Knotsand Physics”, L. Kauffman (ed.), American Mathematical Society, Providence, Rhode Island,1996, [gr-qc/9504036]

[41] N. J. Vilenkin, “Special Functions and the Theory of Group Representations”, American Math-ematical Society, Providence, Rhode Island, 1968

[42] T. Thiemann, O. Winkler, “Gauge Field Theory Coherent States (GCS): IV. Infinite Ten-sor Product and Thermodynamical Limit”, Class. Quantum Grav. 18 (2001) 4997-5033, hep-th/0005235M. Arnsdorf, “Loop Quantum Gravity on Noncompact Spaces”, Nucl. Phys. B577 (2000) 529-546, [gr-qc/9909053]

85

[43] A. Ashtekar, J. Lewandowski, D. Marolf, J. Mourao, T. Thiemann, “Quantization for diffeo-morphism invariant theories of connections with local degrees of freedom”, Journ. Math. Phys.36 (1995) 6456-6493, [gr-qc/9504018]

[44] C. Rovelli, L. Smolin, “Discreteness of volume and area in quantum gravity”, Nucl. Phys. B442(1995) 593, Erratum : Nucl. Phys. B456 (1995) 734 A. Ashtekar, J. Lewandowski, “QuantumTheory of Geometry I: Area Operators”, Class. Quantum Grav. 14 (1997) A55-81

[45] A. Ashtekar, J. Lewandowski, “Quantum Theory of Geometry II : Volume Operators”, Adv.Theo. Math. Phys. 1 (1997) 388-429J. Lewandowski, Class. Quantum Grav. 14 (1997) 71-76R. Loll, “Spectrum of the Volume Operator in Quantum Gravity”, Nucl.Phys.B460:143-154,1996[gr-qc/9511030]R. De Pietri, C. Rovelli, “Geometry Eigenvalues and Scalar Product from Recoupling Theoryin Loop Quantum Gravity”, Phys. Rev. D54 (1996) 2664, [gr-qc/9602023]T. Thiemann, “Closed Formula for the Matrix Elements of the Volume Operator in CanonicalQuantum Gravity”, Journ. Math. Phys. 39 (1998) 3347-3371, [gr-qc/9606091]

[46] T. Thiemann, “A Length Operator for Canonical Quantum Gravity”, Journ. Math. Phys. 39(1998) 3372-3392, [gr-qc/9606092]

[47] T. Thiemann, “Spatially Diffeomorphism Invariant Geometrical Operators in Quantum GeneralRelativity”, work in progress

[48] T. Thiemann, “Anomaly-free Formulation of non-perturbative, four-dimensional LorentzianQuantum Gravity”, Physics Letters B380 (1996) 257-264, [gr-qc/9606088]T. Thiemann, “Quantum Spin Dynamics (QSD): I.”, Class. Quantum Grav. 15 (1998) 839-73,[gr-qc/9606089]; “II. The Kernel of the Wheeler-DeWitt Constraint Operator”, Class. Quan-tum Grav. 15 (1998) 875-905, [gr-qc/9606090]; “III. Quantum Constraint Algebra and PhysicalScalar Product in Quantum General Relativity”, Class. Quantum Grav. 15 (1998) 1207-1247,[gr-qc/9705017]; “IV. 2+1 Euclidean Quantum Gravity as a model to test 3+1 Lorentzian Quan-tum Gravity”, Class. Quantum Grav. 15 (1998) 1249-1280, [gr-qc/9705018]; “V. Quantum Grav-ity as the Natural Regulator of the Hamiltonian Constraint of Matter Quantum Field Theories”,Class. Quantum Grav. 15 (1998) 1281-1314, [gr-qc/9705019]; “VI. Quantum Poincare Algebraand a Quantum Positivity of Energy Theorem for Canonical Quantum Gravity”, Class. Quan-tum Grav. 15 (1998) 1463-1485, [gr-qc/9705020]; “Kinematical Hilbert Spaces for Fermionic andHiggs Quantum Field Theories”, Class. Quantum Grav. 15 (1998) 1487-1512, [gr-qc/9705021]

[49] H. Sahlmann, T. Thiemann, “Towards the QFT on Curved Spacetime Limit of QGR. 1. AGeneral Scheme”, [gr-qc/0207030]; “2. A Concrete Implementation”, [gr-qc/0207031]

[50] D. Marolf, J. Lewandowski, “Loop Constraints : A Habitat and their Algebra”,Int.J.Mod.Phys.D7:299-330,1998, [gr-qc/9710016]R. Gambini, J. Lewandowski, D. Marolf, J. Pullin, “On the Consistency of the ConstraintAlgebra in Spin Network Gravity”, Int.J.Mod.Phys.D7:97-109,1998, [gr-qc/9710018]

[51] M. Gaul, C. Rovelli, “A Generalized Hamiltonian Contraint Operator in Loop Quantum Grav-ity and its Simplest Euclidean Matrix Elements”, Class.Quant.Grav.18:1593-1624,2001, [gr-qc/0011106]

[52] C. Kiefer, Lect.Notes Phys.541:158-187,2000, [gr-qc/9906100] A.O. Barvinsky, “Quantum Cos-mology at the Turn of the Millenium”, [gr-qc/0101046]J. B. Hartle, “Quantum Cosmology: Problems for the 21st Century”, [gr-qc/9701022]

[53] M. Bojowald, “Loop Quantum Cosmology. I. Kinematics”, Class. Quantum Grav. 17(2000) 1489 [gr-qc/9910103]; “II. Volume Operators”, Class. Quantum Grav. 17 (2000)1509 [gr-qc/9910104]; “III. Wheeler-DeWitt Operators”, Class.Quant.Grav.18:1055-1070,2001

86

[gr-qc/0008052]; “IV. Discrete Time Evolution” Class.Quant.Grav.18:1071-1088,2001 [gr-qc/0008053]; “Absence of Singularity in Loop Quantum Cosmology”, Phys.Rev.Lett.86:5227-5230,2001, [gr-qc/0102069]; “Dynamical Initial Conditions in Quantum Cosmology”,Phys.Rev.Lett.87:121301,2001, [gr-qc/0104072]; “The Inverse Scale Factor in Isotropic Quan-tum Geometry”, Phys.Rev.D64:084018,2001, [gr-qc/0105067]; “The Semiclassical Limit of LoopQuantum Cosmology”, Class.Quant.Grav.18:L109-L116,2001, [gr-qc/0105113]; “Inflation fromQuantum Geometry”, gr-qc/0206054

[54] J. C. Baez , “An Introduction to Spin Foam Models of Quantum Gravity and BF Theory”,Lect.Notes Phys.543:25-94,2000, [gr-qc/9905087]J. C. Baez, “Spin Foam Models”, Class.Quant.Grav.15:1827-1858,1998, [gr-qc/9709052]J. W. Barrett, “State Sum Models for Quantum Gravity”, gr-qc/0010050; “Quantum Gravityas Topological Quantum Field Theory”, J.Math.Phys.36:6161-6179,1995, [gr-qc/9506070]A. Perez, “Spin Foam Models for Quantum Gravity”, to appear

[55] M. Reisenberger, C. Rovelli, “Sum over Surfaces Form of Loop Quantum Gravity”, Phys. Rev.D56 (1997) 3490-3508

[56] G. Roepstorff, “Path Integral Approach to Quantum Physics: An Introduction”, Springer Ver-lag, Berlin, 1994

[57] S. Holst, “Barbero’s Hamiltonian Derived From a Generalized Hilbert-Palatini Action”, Phys.Rev. D53 (1996) 5966, [gr-qc/9511026]N. Barros e Sa, “Hamiltonian Analysis of General Relativity with the Immirzi Parameter”, Int.J. Mod. Phys. D10 (2001) 261-272, [gr-qc/0006013]

[58] L. Freidel, K. Krasnov, R. Puzio, “BF Description of Higher Dimensional Gravity Theories”,Adv. Theor. Math. Phys. 3 (1999) 1289-1324, [hep-th/9901069]

[59] J. W. Barrett, L. Crane, “Relativistic Spin Networks and Quantum Gravity”, J. Math. Phys.39 (1998) 3296-3302, [gr-qc/9709028]

[60] J. C. Baez, J. W. Barrett, “Integrability of Relativistic Spin Networks”, gr-qc/0101107A. Perez, C. Rovelli, “Spin Foam Model for Lorentzian General Relativity”, Phys. Rev. D63(2001) 041501, [gr-qc/0009021]L. Crane, A. Perez, C. Rovelli, “A Finiteness Proof for the Lorentzian State Sum Spin FoamModel for Quantum General Relativity”, gr-qc/0104057; “Perturbative Finiteness in Spin-FoamQuantum Gravity”, Phys. Rev. Lett. 87 (2001) 181301

[61] D. V. Boulatov, Mod. Phys. Lett. A7 (1992) 1629H. Ooguri, Mod. Phys. Lett. A7 (1992) 2799

[62] F. Markopoulou, L. Smolin, “Causal Evolution of Spin Networks”, Nucl.Phys.B508:409-430,1997, [gr-qc/9702025]F. Markopoulou, “Dual Formulation of Spin Network Evolution”, gr-qc/9704013F. Markopoulou, L. Smolin, “Quantum Geometry with Intrinsic Local Causality”,Phys.Rev.D58:084032,1998, [gr-qc/9712067]F. Markopoulou, “The Internal Description of a Causal Set: What the Universe Like fromInside”, Commun.Math.Phys.211:559-583,2000, [gr-qc/9811053]F. Markopoulou, “Quantum Causal Histories”, Class.Quant.Grav.17:2059-2072,2000, [hep-th/9904009]F. Markopoulou, “An Insider’s Guide to Quantum Causal Histories”,Nucl.Phys.Proc.Suppl.88:308-313,2000, [hep-th/9912137]

[63] F. Markopoulou, “An Algebraic Approach to Coarse Graining”, hep-th/0006199[64] J. C. Baez, J. D. Christensen, T. R. Halford, D. C. Tsang, “Spin Foam Models of Riemannian

Quantum Gravity”, Class. Quant. Grav. 19 (2002) 4627-4648, [gr-qc/0202017]

87

A. Perez, “Spin Foam Quantization of Plebanski’s Action”, Adv. Theor. Math. Phys. 5 (2002)947-968, [gr-qc/0203058]

[65] M. Bojowald and A. Perez, “Spin Foam Quantization and Anomalies”, in preparation[66] J. D. Bekenstein, “Black Holes and Entropy”, Phys.Rev. D7 (1973) 2333-2346

J. D. Bekenstein, “Generalized Second Law for Thermodynamics in Black Hole Physics”, Phys.Rev. D9 (1974) 3292-3300S.W. Hawking, “Particle Creation by Black Holes”, Commun. Math. Phys. 43 (1975) 199-220

[67] K. Krasnov, “On Statistical Mechanics of Gravitational Systems”, Gen. Rel. Grav. 30 (1998)53-68, [gr-qc/9605047]C. Rovelli, “Black Hole Entropy from Loop Quantum Gravity” Phys. Rev. Lett. 77 (1996)3288-3291, [gr-qc/9603063]

[68] A. Ashtekar, C. Beetle, O. Dreyer, S. Fairhurst, B. Krishnan, J. Lewandowski, J. Wisniewski“Isolated Horizons and Their Applications”, Phys.Rev.Lett.85:3564-3567,2000 [gr-qc/0006006]A. Ashtekar, “Classical and Quantum Physics of Isolated Horizons”, Lect.Notes Phys.541:50-70,2000A. Ashtekar, “Interface of general Relativity, Quantum Physics and Statistical Mecahnics: Somerecent Developments”, Annalen Phys.9:178-198,2000, [gr-qc/9910101]

[69] A. Ashtekar, A. Corichi, K. Krasnov, “Isolated Horizons: The Classical Phase Space”,Adv.Theor.Math.Phys.3:419-478,2000, [gr-qc/9905089]A. Ashtekar, J. C. Baez, K. Krasnov, “Quantum Geometry of Isolated Horizons and Black HoleEntropy”, Adv.Theor.Math.Phys.4:1-94,2001, [gr-qc/0005126]

[70] T. Regge, C. Teitelboim, “Role of Surface Integrals in the Hamiltonian Formulation of GeneralRelativity”, Annals Phys. 88 (1974) 286

[71] L. Smolin, “Linking Topological Quantum Field Theory and Non-Perturbative Quantum Grav-ity”, J. Math. Phys. 36 (1995) 6417, [gr-qc/9505028]

[72] S. Axelrod, S. D. Pietra, E. Witten, “Geometric Quantization of Chern-Simons Gauge Theory”,J. Diff. Geo. 33 (1991) 787-902

[73] G. ’t Hooft, “The Holographic Principle: Opening Lecture”, in “Erice 1999, Basics and highlightsin fundamental physics”, 397-413, [hep-th/0003004]

[74] T. Thiemann, “Reality Conditions inducing Transforms for Quantum Gauge Field Theo-ries and Quantum Gravity”, Class. Quantum Gravity 13 (1996) 1383-1403, [gr-qc/9511057];“An Account of Transforms on A/G”, Acta Cosmologica 21 (1995) 145-167, [gr-qc/9511049];“Gauge Field Theory Coherent States (GCS): I. General Properties”, Class.Quant.Grav.18:2025-2064,2001, [hep-th/0005233]; “Complexifier Coherent States for Quantum General Relativity”,[gr-qc/0206037]

[75] T. Thiemann, “Quantum Spin Dynamics (QSD): VII. Symplectic Structures and ContinuumLattice Formulations of Gauge Field Theories”, Class.Quant.Grav.18:3293-3338,2001, [hep-th/0005232]T. Thiemann, O. Winkler, “Gauge Field Theory Coherent States (GCS): II. Peakedness Proper-ties”, Class.Quant.Grav.18:2561-2636,2001, [hep-th/0005237]; “III. Ehrenfest Theorems”, Class.Quantum Grav. 18 (2001) 4629-4681, [hep-th/0005234]H. Sahlmann, T. Thiemann, O. Winkler, “ Coherent States for Canonical Quantum Gen-eral Relativity and the Infinite Tensor product Extension”, Nucl.Phys.B606:401-440,2001, [gr-qc/0102038]

[76] B. C. Hall, Journ. Funct. Analysis 122 (1994) 103B. C. Hall, J. J. Mitchell, “Coherent States on Spheres”, J. Math. Phys. 43 (2002) 1211-1236, [quant-ph/0109086]; “The Large Radius Limit for Coherent States on Spheres”, quant-ph/0203142

88

[77] G. Amelino-Camelia, “Are we at Dawn with Quantum Gravity Phenomenology”, Lectures givenat 35th Winter School of Theoretical Physics: From Cosmology to Quantum Gravity, Polanica,Poland, 2-12 Feb 1999, Lect.Notes Phys.541:1-49,2000, [gr-qc/9910089]G. Amelino-Camelia, John R. Ellis, N.E. Mavromatos, D.V. Nanopoulos, Subir Sarkar, “Po-tential Sensitivity of Gamma Ray Burster Observations to Wave Dispersion in Vacuo”, Nature393:763-765,1998, [astro-ph/9712103] R. Gambini, J. Pullin, “Nonstandard Optics from Quan-tum Spacetime” Phys.Rev.D59:124021,1999, [gr-qc/9809038]R. Gambini, J. Pullin, “Quantum Gravity Experimental Physics ?”, Gen.Rel.Grav.31:1631-1637,1999J. Alfaro, H. A. Morales-Tecotl, L. F. Urrutia, “Quantum Gravity Corrections to Neutrino Prop-agation”, Phys.Rev.Lett.84:2318-2321,2000, [gr-qc/9909079]J. Alfaro, H. A. Morales-Tecotl, L. F. Urrutia, “Loop Quantum Gravity and Light Propagation”,Phys. Rev. D65 (2002) 103509, [hep-th/0108061]

[78] S. D. Biller et al, Phys. Rev. Lett. 83 (1999) 2108[79] A. Ashtekar, Carlo Rovelli, L. Smolin, “Gravitons and Loops”, Phys.Rev. D44 (1991) 1740-1755,

[hep-th/9202054][80] M. Varadarajan, “Fock representations from U(1) Holonomy Algebras”, Phys. Rev. D61 (2000)

104001 [gr-qc/0001050]; “Photons from Quantized Electric Flux Representations”, Phys.Rev.D64 (2001) 104003, [gr-qc/0104051]; M. Varadarajan, “Gravitons from a Loop Representationof Linearized Gravity”, Phys. Rev. D66 (2002) 024017, [gr-qc/0204067]

[81] J. Velhinho, “Invariance Properties of Induced Fock Measures for U(1) Holonomies”, Commun.Math. Phys. 227 (2002) 541-550, [math-ph/0107002]

[82] L. Smolin, “Strings as Perturbations of Evolving Spin Networks”, Nucl.Phys.Proc.Suppl.88:103-113,2000, [hep-th/9801022]; “A Holographic Formulation of Quantum General Relativity”,Phys.Rev.D61:084007,2000, [hep-th/9808191]; “Towards a Background Independent Approachto M Theory”, hep-th/9808192; “The Cubic Matrix Model and Duality Between Strings andLoops”, [hep-th/0006137]; “A Candidate for a Background Independent Formulation of M The-ory”, Phys.Rev.D62:086001,2000, [hep-th/9903166]; “The Exceptional Jordan Algebra and theMatrix String”, hep-th/0104050Y. Ling, L. Smolin, “Eleven – Dimensional Supergravity as a Constrained Topological FieldTheory”, Nucl.Phys.B601:191-208,2001, [hep-th/0003285]; “Supersymmetric Spin Networks andQuantum Supergravity”, Phys.Rev.D61:044008,2000, [hep-th/9904016]; “Holographic Formula-tion of Quantum Supergravity”, Phys.Rev.D63:064010,2001, [hep-th/0009018]

[83] L. Smolin, “M Theory as a Matrix Extension of Chern-Simons Theory”, Nucl.Phys.B591:227-242,2000, [hep-th/0002009]; L. Smolin, “Quantum Gravity with a Positive Cosmological Con-stant”, hep-th/0209079

[84] R. Helling, H. Nicolai, “Supermebranes and Matrix Theory”, hep-th/9809103[85] K. Pohlmeyer, “A Group Theoretical Approach to the Quantization of the Free Relativistic

Closed String”, Phys. Lett. B119 (1982) 100K. Pohlmeyer, K.H. Rehren, “Algebraic Properties of the Invariant Charges of the Nambu-GotoTheory”, Commun.Math.Phys. 105 (1986) 593; “The Algebra formed by the Charges of theNambu-Goto Theory: Identification of a Maximal Abelean Subalgebra” Commun. Math. Phys.114 (118) 55; “The Algebra formed by the Charges of the Nambu-Goto Theory: Their Geomet-ric Origin and Their Completeness”, Commun. Math. Phys. 114 (1988) 177K. Pohlmeyer, “The Invariant Charges of the Nambu-Goto Theory in WKB Approximation”,Commun. Math. Phys. 105 (1986) 629; “The Algebra formed by the Charges of the Nambu-Goto Theory: Casimir Elements”, Commun. Math. Phys. 114 (1988) 351; “Uncovering the

89

Detailed Structure of the Algebra Formed by the Invariant Charges of Closed Bosonic StringsMoving in (1+2)-Dimensional Minkowski Space”, Commun. Math. Phys. 163 (1994) 629-644;“The Invariant Charges of the Nambu-Goto Theory: Non-Additive Composition Laws”, Mod.Phys. Lett. A10 (1995) 295-308; “The Nambu-Goto Theory of Closed Bosonic Strings Movingin (1+3)-Dimensional Minkowski Space: The Quantum Algebra of Observables”, Annalen Phys.8 (1999) 19-50, [hep-th/9805057]K. Pohlmeyer, M. Trunk, “The Invariant Charges of the Nambu-Goto Theory: Quantization ofNon-Additive Composition Laws”, hep-th/0206061G. Handrich, C. Nowak, “The Nambu-Goto Theory of Closed Bosonic Strings Moving in (1+3)-Dimensional Minkowski Space: The Construction of the Quantum Algebra of Observables upto Degree Five”, Annalen Phys. 8 (1999) 51-54, [hep-th/9807231]G. Handrich, “Lorentz Covariance of the Quantum Algebra of Observables: Nambu-Goto Stringsin 3+1 Dimensions”, Int. J. Mod. Phys. A17 (2002) 2331-2349G. Handrich, C. Paufler, J.B. Tausk, M. Walter, “The Representation of the Algebra of Ob-servables of the Closed Bosonic String in 1+3 Dimensions: Calculation to Order ~7”, math-ph/0210024C. Meusburger, K.H. Rehren, “Algebraic Quantization of the Closed Bosonic String”, math-ph/0202041

90